Hongik - Ultrathin Organic Films Grown by Organic Molecular...

104
Ultrathin Organic Films Grown by Organic Molecular Beam Deposition and Related Techniques Stephen R. Forrest Center for Photonics and Optoelectronic Materials (POEM), Department of Electrical Engineering and the Princeton Materials Institute, Princeton University, Princeton, New Jersey 08544 Received May 2, 1997 (Revised Manuscript Received July 9, 1997) Contents 1.0. Introduction 1793 2.0. Techniques for UHV Growth of Organic Thin Films 1795 2.1. Purification of Source Materials 1796 2.2. Maintaining Purity during Film Growth 1797 3.0. Structure of Organic Thin Films Grown by OMBD 1799 3.1. Definition of Growth Modes 1799 3.2. Epitaxy and van der Waals Epitaxy 1800 3.3. Quasi-Epitaxy 1805 3.3.1. Theory of QE 1805 3.3.2. Observation of QE Growth 1816 3.4. Growth of Other Multilayer Structures By OMBD 1837 3.4.1. Organic/Organic Multilayer Structures 1837 3.4.2. Organic/Inorganic Hybrid Multilayer Structures 1840 3.5. Summary of Growth Results 1841 4.0. Optoelectronic Properties of OMBD-Grown Films 1842 4.1. Excitons in Ultrathin Organic Films Grown by OMBD 1842 4.1.1. Spectroscopic Identification of Excited States in PTCDA 1843 4.1.2. Fluorescence and Absorption Spectra of PTCDA/NTCDA MQWs 1845 4.2. CT States in Nonpolar OMCs: Theory and Experimental Confirmation 1849 4.2.1. Hamiltonian for CT Excitons 1850 4.2.2. Theoretical Fits to Electroabsorption and MQW Absorption Spectra 1853 4.2.3. Correspondence between MQW Absorption and Fluorescence Spectra 1857 4.3. Optical and Electronic Properties of Other Organic Multilayer Systems 1858 4.4. Implications and Applications of Exciton Confinement in OMCs 1861 5.0. Applications of Thin-Film Structures Grown in Vacuum 1862 5.1. Integrated Organic-on-Inorganic Heterojunction Devices 1863 5.2. Organic Solar Cells 1866 5.3. Molecular Organic Light-Emitting Devices 1871 5.3.1. Current Transport and Electroluminescence Mechanisms in OLEDs 1873 5.3.2. Organic Light-Emitting Device Structures and Applications 1877 5.4. Thin-Film Transistors 1880 6.0. The Future of Ordered Organic Thin Films Grown in the Vapor Phase 1883 6.1. Limitations of OMBD 1883 6.2. Applications on the Horizon 1888 7.0. Conclusions 1890 8.0. Acknowledgments 1891 9.0. References 1891 1.0. Introduction During the past decade, enormous progress has been made in growing ultrathin organic films and multilayer structures with a wide range of exciting optoelectronic properties. This progress has been made possible by several important advances in our understanding of organic films and their modes of growth. Perhaps the single most important advance has been the use of ultrahigh vacuum (UHV) as a means to achieve, for the first time, monolayer control over the growth of organic thin films with extremely high chemical purity and structural precision. 1-3 Such monolayer control has been pos- sible for many years using well-known techniques such as Langmuir-Blodgett film deposition, 4 and more recently, self-assembled monolayers from solu- tion have also been achieved. 5 However, ultrahigh- vacuum growth, sometimes referred to as organic molecular beam deposition (OMBD) or organic molecular beam epitaxy (OMBE), has the advantage of providing both layer thickness control and an atomically clean environment and substrate. When combined with the ability to perform in situ high- resolution structural diagnostics of the films as they are being deposited, techniques such as OMBD have provided an entirely new prospect for understanding many of the fundamental structural and optoelec- tronic properties of ultrathin organic film systems. Since such systems are both of intrinsic as well as practical interest, substantial effort worldwide has been invested in attempting to grow and investigate the properties of such thin-film systems. This paper is a review of recent progress made in organic thin films grown in ultrahigh vacuum or using other vapor-phase deposition methods. We will describe the most important work which has been published in this field since the emergence of OMBD in the mid-1980s. Both the nature of thin-film growth and structural ordering will be discussed, as well as some of the more interesting consequences to the physical properties of such organic thin-film systems will be considered both from a theoretical as well as an experimental viewpoint. Indeed, it will 1793 Chem. Rev. 1997, 97, 1793-1896 S0009-2665(94)01014-9 CCC: $28.00 © 1997 American Chemical Society

Transcript of Hongik - Ultrathin Organic Films Grown by Organic Molecular...

Page 1: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

Ultrathin Organic Films Grown by Organic Molecular Beam Deposition andRelated Techniques

Stephen R. Forrest

Center for Photonics and Optoelectronic Materials (POEM), Department of Electrical Engineering and the Princeton Materials Institute,Princeton University, Princeton, New Jersey 08544

Received May 2, 1997 (Revised Manuscript Received July 9, 1997)

Contents

1.0. Introduction 17932.0. Techniques for UHV Growth of Organic Thin

Films1795

2.1. Purification of Source Materials 17962.2. Maintaining Purity during Film Growth 1797

3.0. Structure of Organic Thin Films Grown by OMBD 17993.1. Definition of Growth Modes 17993.2. Epitaxy and van der Waals Epitaxy 18003.3. Quasi-Epitaxy 18053.3.1. Theory of QE 18053.3.2. Observation of QE Growth 1816

3.4. Growth of Other Multilayer Structures ByOMBD

1837

3.4.1. Organic/Organic Multilayer Structures 18373.4.2. Organic/Inorganic Hybrid Multilayer

Structures1840

3.5. Summary of Growth Results 18414.0. Optoelectronic Properties of OMBD-Grown Films 1842

4.1. Excitons in Ultrathin Organic Films Grown byOMBD

1842

4.1.1. Spectroscopic Identification of ExcitedStates in PTCDA

1843

4.1.2. Fluorescence and Absorption Spectra ofPTCDA/NTCDA MQWs

1845

4.2. CT States in Nonpolar OMCs: Theory andExperimental Confirmation

1849

4.2.1. Hamiltonian for CT Excitons 18504.2.2. Theoretical Fits to Electroabsorption and

MQW Absorption Spectra1853

4.2.3. Correspondence between MQWAbsorption and Fluorescence Spectra

1857

4.3. Optical and Electronic Properties of OtherOrganic Multilayer Systems

1858

4.4. Implications and Applications of ExcitonConfinement in OMCs

1861

5.0. Applications of Thin-Film Structures Grown inVacuum

1862

5.1. Integrated Organic-on-InorganicHeterojunction Devices

1863

5.2. Organic Solar Cells 18665.3. Molecular Organic Light-Emitting Devices 18715.3.1. Current Transport and

Electroluminescence Mechanisms inOLEDs

1873

5.3.2. Organic Light-Emitting Device Structuresand Applications

1877

5.4. Thin-Film Transistors 1880

6.0. The Future of Ordered Organic Thin FilmsGrown in the Vapor Phase

1883

6.1. Limitations of OMBD 18836.2. Applications on the Horizon 1888

7.0. Conclusions 18908.0. Acknowledgments 18919.0. References 1891

1.0. IntroductionDuring the past decade, enormous progress has

been made in growing ultrathin organic films andmultilayer structures with a wide range of excitingoptoelectronic properties. This progress has beenmade possible by several important advances in ourunderstanding of organic films and their modes ofgrowth. Perhaps the single most important advancehas been the use of ultrahigh vacuum (UHV) as ameans to achieve, for the first time, monolayercontrol over the growth of organic thin films withextremely high chemical purity and structuralprecision.1-3 Such monolayer control has been pos-sible for many years using well-known techniquessuch as Langmuir-Blodgett film deposition,4 andmore recently, self-assembled monolayers from solu-tion have also been achieved.5 However, ultrahigh-vacuum growth, sometimes referred to as organicmolecular beam deposition (OMBD) or organicmolecular beam epitaxy (OMBE), has the advantageof providing both layer thickness control and anatomically clean environment and substrate. Whencombined with the ability to perform in situ high-resolution structural diagnostics of the films as theyare being deposited, techniques such as OMBD haveprovided an entirely new prospect for understandingmany of the fundamental structural and optoelec-tronic properties of ultrathin organic film systems.Since such systems are both of intrinsic as well aspractical interest, substantial effort worldwide hasbeen invested in attempting to grow and investigatethe properties of such thin-film systems.This paper is a review of recent progress made in

organic thin films grown in ultrahigh vacuum orusing other vapor-phase deposition methods. We willdescribe the most important work which has beenpublished in this field since the emergence of OMBDin the mid-1980s. Both the nature of thin-filmgrowth and structural ordering will be discussed, aswell as some of the more interesting consequencesto the physical properties of such organic thin-filmsystems will be considered both from a theoreticalas well as an experimental viewpoint. Indeed, it will

1793Chem. Rev. 1997, 97, 1793−1896

S0009-2665(94)01014-9 CCC: $28.00 © 1997 American Chemical Society

Page 2: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

be a major objective of this review to place theimportance of ultrahigh vacuum deposition in con-text. As of this writing, there remains considerablediscussion as to the necessity for the use of ultrahighvacuum, and the implication of the need for ultrahighpurity. For example, the attraction of organics formany applications often hinges on the belief thatmaterial purity is not important, thus allowing forthe realization of devices which are comparativelysimple to fabricate and process. These argumentswill be considered, and we will attempt to addresstheir importance in various applications.Beyond the use of UHV to achieve extreme chemi-

cal and structural control over organic thin films, itis useful to consider the ultimate motivations forembarking on such a difficult and expensive researchendeavor. Perhaps the most compelling reason forstudying organic nanostructures stems from thepromising scientific and technological results ofsimilar research performed over a span of 25 yearsin conventional semiconductor heterojunctions grownby the analogous UHV process of molecular beamepitaxy (MBE).6 By controlling inorganic semicon-ductor thin film thickness and structure, a widerange of useful and interesting devices based onmultiple quantum wells (MQWs) have been demon-strated including low-threshold current laser diodes,low-noise avalanche photodetectors, and high-band-

width optical modulators.7-11 The advantageousproperties of these MQW structures stems from twofactors controlled by the precision of the epitaxialgrowth process employed: the density of conductionand valence band states (leading to low laser thresh-old currents12 and the quantum confined Stark ef-fect13 employed in optical modulators), and theheterojunction energy band offsets between contact-ing materials of different composition (leading to thecontrol of current transport). Considerable work inorganic thin films has likewise been directed towardachieving these same goals, striving to tailor thematerials properties via control of the densities ofstates and the energy band offsets between differingmaterials. In addition to these objectives, organicmaterials in their infinite variety have also beeninvestigated for their application to nonlinearoptics14-18 and other optoelectronic devices.19-22 In-deed, progress in the growth of organic thin-filmnanostructures by UHV techniques such as thedemonstration23,24 of organic MQWs has spawnedconsiderable theoretical work, predicting completelynew nonlinear optical phenomena in organic/organicand organic/inorganic MQW structures.25-29 Hence,we find ourselves at the threshold of realizing anentirely new class of materials, offering nearly un-precedented opportunities for extending our under-standing of the fundamental properties of a largeclass of materialssnanostructured organic molecularcrystals. These structures, in turn, promise to havemany new and useful properties which will undoubt-edly be heavily exploited well into the next century.Even at this relatively early stage of development,vacuum-deposited heterojunction “small molecule”-based organic light-emitting devices are being com-mercially introduced as the light sources in novel,bright, flat panel displays.30-32 The applications forsuch vacuum-deposited organic materials is limitedmainly by our ability to control composition andstructure of the resulting thin films.The study of vapor-deposited organic films has

been applied to a vast array of molecular systems.However, most of the work which has made thegreatest impact to date has concentrated on the studyof the growth and optoelectronic characteristics ofplanar stacking molecules such as the phthalocya-nines,18,21,33,34 and polycyclic aromatic compoundsbased on naphthalene and perylene. In particular,a molecular system based on perylenesthe planarstacking 3,4,9,10-perylenetetracarboxylic dianhy-dride (PTCDA)shas become the focus of considerableattention in both our own laboratory as well as manylaboratories worldwide.35-43 For this reason, muchof the discussion will focus on the growth andproperties of the Pc’s and aromatic compounds basedon naphthalene and perylene which can be treatedas archetypes of a larger class of planar stackingmolecules which behave in analogous, although notnecessarily identical, ways to each other. In addition,some recent results with more unusual materialssuch as organic salts, and other highly polar com-pounds will be discussed.This review is organized as follows: In section 2,

the technology of ultrahigh-vacuum growth of organicthin films will be discussed, followed by section 3

Stephen Forrest graduated from the University of California with a B.A.degree in physics in 1972, and from the University of Michigan with aM.Sc. and Ph.D. degree in physics in1974 and 1979, respectively. Fromthere, he went to Bell Laboratories (Murray Hill) where he did bothfundamental and applied research and development of photodetectorsfor use in long wavelength optical communications systems. In 1982, hebecame supervisor of the Integrated Optoelectronic Devices and Circuitsgroup at Bell Laboratories. There, he worked on arrays of emitters anddetectors, and integrated optical receivers. In 1985, Professor Forrestjoined the faculty of the Departments of Electrical Engineering andMaterials Science at the University of Southern California where hecontinued his research on optoelectronic integrated circuits, as well ason a new class of optoelectronic materials: crystalline organic semicon-ductors. From 1989 to 1992, Professor Forrest served as the Director ofthe National Center for Integrated Photonic Technology: a consortium offive universities including USC, Columbia, Kent State University, MTT,and UCLA. In 1992, Professor Forrest joined Princeton University asthe James S. McDonnell Distinguished University Professor of ElectricalEngineering and the Princeton Materials Institute, and as Director ofPrinceton’s Advanced Technology Center for Photonics and OptoelectronicMaterials (ATC/POEM). Professor Forrest has served as Associate Editorto the Journal of Quantum Electronics and Photonics Technology Letters,has served on the OSA Technical Council, and is on the LEOS Board ofGovernors. In 1996−97, he was the recipient of the IEEE/LEOSDistinguished Lecturer Award. He is a member of the APS, MRS, andthe OSA, and is a Fellow of the IEEE.

1794 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 3: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

where the growth modes of organic thin films willbe discussed. We will place the various forms ofepitaxy, in particular van der Waals epitaxy and thenewly discovered quasi-epitaxy in the context ofconventional epitaxy as it applies to inorganic crys-talline systems. The relative importance of purityand structural control will also be considered in thissection. In section 4, we will review recent resultson the optical and electronic properties of ultrathinorganic films grown from the vapor phase. In par-ticular, exciton confinement and charge transfer inultrathin multilayer stacks will be discussed fromboth theoretical as well as experimental viewpoints.In section 5 we will consider several emerging ap-plications for vapor-deposited organic thin films suchas for nonlinear optical and electroluminescent de-vices. Finally, in section 6 we will summarize therecent findings in the field, and discuss the futureprospects of vapor-deposited thin organic films bothas they relate to improving our fundamental under-standing of these condensed matter systems, as wellas to their prospects for use in advanced deviceapplications.

2.0. Techniques for UHV Growth of Organic ThinFilmsThe most common means for the OMBD growth of

organic thin films is to use an ultrahigh-vacuumapparatus such as that shown in Figure 2-1 whichis similar, in many respects, to conventional molec-ular beam epitaxial growth systems. Typically,growth occurs by the evaporation in a backgroundvacuum ranging from 10-7 to 10-11 Torr of a highlypurified powder or, less frequently, a melt of theorganic source material from a temperature-con-trolled oven or Knudsen cell.6 (It is customary toquote pressures in UHV environments using unitsof Torr, where 1 Torr ) 1 mmHg, or in SI units ofPascal, where 1 Pa ) 1 N/m2. The conversionbetween these conventions is 1 Pa ) 7.5 × 10-3 Torr.For this review, all pressures will be given in unitsof Torr.) The evaporant is collimated by passingthrough a series of orifices, after which it is depositedon a substrate held perpendicular to the beamapproximately 10-20 cm from the source. The fluxof the molecular beam is controlled by a combinationof oven temperature (which is maintained somewhatabove the sublimation temperature of the source, andwell below its chemical decomposition temperature),as well as a mechanical shutter which can switch thebeam flux from “on” to “off”. By sequentially shut-tering the beam flux from more than one Knudsencell, multilayer structures consisting of alternatinglayers of different compounds can be grown by thismethod. Indeed, using sequential shuttering, organicmultilayer structures such as multiple quantumwells have been demonstrated.23,24Typical growth rates range from 0.001 to 100 Å/s.

Since the monolayer (ML) thickness is from 3 to 5 Å,these rates correspond to from 0.7 ML/h to 30 ML/s.The lower end of this range has a danger of adsorbingcontaminants onto the surface at a higher rate thanthe organic molecules, whereas at the higher rates,growth is extremely difficult to control. For thesereasons, most reliable growth data have been ob-

tained at rates of from 0.1 to 5 Å/s. Growth rate ismeasured using a quartz crystal thickness monitorof the type used in metal thin-film deposition. Be-tween growth runs, the Knudsen cell is often main-tained at somewhat below the material sublimationtemperature in order to continuously outgas thesource. In this manner, very high purity sourcematerials can be achieved and maintained over longperiods of time. We speculate that the UHV storageof the source material at elevated temperatures is akey factor in eliminating impurities and moisturewhich lead to film defects during growth.Substrate temperature is also an important vari-

able in determining film structure.1,44-49 A typicalrange of growth temperatures is from 80 to 400 K.Note that at lower temperatures, impurities tend toquickly adsorb onto the substrate. This is particu-larly important when growth occurs at only modestbackground pressures. Under almost all substratetemperatures employed for organic film growth, weassume that the molecules have a sticking coefficientequal to unitysthat is, all molecules striking thesubstrate surface adhere to that surface. This situ-ation results since the molecules are large and thesubstrate temperature is maintained well below thedesorption temperature. However, some researchershave used substrate temperature control to tune thedesorption rate, thereby enabling controlled growthof monolayers of some organic molecular materials.50

One significant advantage of using UHV is thatnumerous in situ film diagnostics can be employedboth during and immediately after growth.6,51 Manygrowth chambers are equipped with reflection highenergy electron diffraction (RHEED) to allow for insitu determination of the thin-film crystalline struc-ture.52,53 In addition, residual gas analysis (RGA) canbe employed to determine the molecular specieswithin the chamber during growth, as well as todetermine the energy of film-substrate adhesionusing thermal desorption spectroscopy (TDS).54 Otherdiagnostics include low-energy electron diffraction55,56(LEED), which typically is done in a chamber at-tached to the growth system due to the incompatiblegeometries of the growth and LEED equipment.Finally, many other vacuum analysis techniques suchas UHV scanning tunneling microscopy57,58 (STM),Auger electron spectroscopy, in situ ellipsometry, etc.have all been employed in conjunction with OMBDsimply by installing the appropriate equipment in achamber which is connected via a vacuum feedthroughto the growth chamber.Hence, OMBD growth of thin organic films has lead

to significant advances in our understanding ofgrowth dynamics as well as fundamental optoelec-tronic properties of ultrathin organic molecular crys-tals (OMCs) due to the high control afforded over thegrown structure, the assurance of high materialpurity, and the ability to incorporate many powerful,in situ diagnostics into the growth system. Many ofthese same advantages have been used with greateffectiveness in growing extremely high quality in-organic semiconductors such as GaAs and InP usingmolecular beam epitaxial techniques. It is not un-reasonable to assume that similar advantages canstem from the use of UHV techniques applied to the

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1795

Page 4: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

growth of OMC thin films, particularly since, as wewill show, many of the constraints of lattice matchingwhich have limited the materials combinations ac-cessible in inorganic semiconductor systems aresignificantly relaxed in the case of organic thin films.Of course, the nearly infinite variety of molecularcompounds available also make OMBD growth ofsuch films a very promising area of investigation.

2.1. Purification of Source MaterialsPurification of source materials is essential for

assuring that the grown thin film is reasonably freeof impurities.59,60 Furthermore, purification is re-

quired to prevent contaminants from entering thehigh vacuum chamber which might result in a highbackground pressure, as well as a constant source ofcontamination of the subsequently grown films dueto outgassing from the deposits in the chamber itself.There are several techniques for purification, includ-ing gradient sublimation,59 zone refining from themelt,61 chromatography,59,61 etc. Although the high-est purity organics have been achieved via zonerefining,61,62 gradient sublimation is the most usefulmeans for purification of the powders employed inOMBD since most of these compounds do not have aliquid phase at atmospheric pressure or below.1,63

Figure 2-1. (a) Cross-sectional view of the organic molecular beam deposition chamber. The rotating substrate holdercan be temperature controlled between 80 and 900 K using a combination of liquid nitrogen cooling and boron nitrideheating elements. (b) Layout of the combined OMBD and gas source MBE system in the author’s laboratory. There aretwo chambers for organic growth (OMBD-I and -II), and the MBE chamber is used for the growth of InP-based materialsand organic/inorganic heterojunctions. Other chambers include sputtering and electron beam deposition for contactfabrication, and an in situ analysis chamber.

1796 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 5: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

Gradient (or train) sublimation purification pro-ceeds as follows: The source materials (such asPTCDA or Copper phthalocyanine, CuPc) as obtainedfrom a commercial source or from laboratory synthe-sis are successively purified prior to loading thesource materials into the ultrahigh vacuum deposi-tion system. Typically, several grams of powderedsource material are loaded into the end of a 60 cmlong glass tube. A glass sleeve is then placed insidethe tube to capture the sublimed material, followedby a wad of glass wool placed at the open end of thetube to prevent the source material from contaminat-ing the pumping chamber (see Figure 2-2). The tubeis evacuated to <1 × 10-6 Torr using a turbomolecu-lar or other dry pump, and the end containing theorganic source material is inserted into a furnace.The temperature of the furnace is gradually in-creased until the sublimation point of the materialis reached (approximately 400 °C for PTCDA, and300 °C for CuPc); a process typically taking severaldays. Purified organic crystals grow on the insidewall of the glass sleeve in the warm zone of thefurnace, while more volatile impurities are evacu-ated. Nonvolatile impurities are left at the hottestend of the glass tube. The glass sleeve insert is thenremoved, and the crystals are extracted to be usedas the source material for the second purificationcycle. Before loading the source materials into theOMBD system, organic compounds are often purifiedin multiple cycles using this method. By the end ofthe process, ∼50% of the original source material hasbeen removed, and the background pressure of thesublimation system is ∼10-7 Torr. The purifiedmaterials can then be loaded immediately into theOMBD effusion cells where they are continuouslymaintained in UHV at elevated temperatures await-ing growth.

2.2. Maintaining Purity during Film GrowthGrowth in a UHV environment assures that a

minimal density of impurities and defects will beincorporated into the film. To our knowledge, thereis no other form of growth of thin films which canassure the low levels of contamination inherent in

the OMBD process. However, as noted above, thedegree to which impurities are adsorbed onto thesubstrate surface depend on the quality of thevacuum. Hence, the impurity incorporation rateshould be substantially less than the film growthrate. By using simple kinetic theory to estimate theimpurity adsorption rate, the time for a monolayerof background gas atoms of mass, m, to be adsorbedonto a substrate surface is64:

where Ns is the density of surface atoms required toform a complete monolayer, ú is the atomic stickingcoefficient, P is the background pressure, k is Bolt-zmann’s constant, and T is the temperature. Assum-ing that the principal residual gas in a vacuumsystem is nitrogen ú ) 1, and forNs ) 1014 molecules/cm2, then the time required to adsorb a monolayeris τ ∼ 3 min at a background pressure of 5 × 10-9

Torr, and 30 min at 5 × 10-10 Torr. Hence, tominimize the incorporation of impurities even at veryhigh vacuums, it is important to employ the fastestgrowth rate possible which is compatible with achiev-ing the desired film structure and thickness control.Practical growth rates thus lie in the range of 0.1 to5 ML/s, which is similar to that used in the MBEgrowth of inorganic semiconductors .We note that the effects of impurities in organic

thin films are not equivalent to their effects inconventional semiconductors such as Si or GaAs. Inthese latter materials, impurities often take the formof interstitial or substitutional lattice defects whichprovide excess carriers via the formation of energylevels lying near to either the valence or conductionband edges. Carriers trapped in these “midgap”states are easily thermalized and hence stronglyaffect the conductivity of the host semiconductor.Most OMCs grown by OMBD, however, are bondedby van der Waals (vdW) forces, and hence impuritiesdo not typically share valences with crystallineenergy bands. The defects, therefore, are generallyinterstitial, with very tight binding of the excesscarriers at these localized sites. However, we point

Figure 2-2. Schematic diagram of the apparatus used in thermal gradient sublimation of organic materials as a meansto achieve high purity prior to OMBD growth (from ref 450).

τ ) Nsú(2πmkT)1/2/P (2.1)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1797

Page 6: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

out that some interstitial defects can be electricallyactive. A well-known example is O2

- in CuPc, wherethe O2 forms a complex with the Cu thereby providinga high density of holes to the crystal.19More typically, impurities in organic thin films

disrupt the regular stacking habits of the molecules.These point defects can thus significantly reduce thecharge carrier mobility, as has been observed invacuum deposited films of PTCDA. This is apparentin Figure 2-3, where charge mobility is plotted as afunction of deposition rate65 (Figure 2-3a) and sub-strate temperature (Figure 2-3b).1 It has previouslybeen found that at room temperature, high depositionrates of this particular material can lead to a higherdegree of stacking order, as does a lower substratetemperature (see section 3.3.2). Hence, we can inferfrom these data that stacking faults in closely spacedplanar molecular systems such as PTCDA can leadto a reduction in carrier mobility. Impurity incor-poration leading to film disorder can also havesimilar effects on the resulting materials properties.In contrast, individual atomic impurities may not

be sufficiently large to induce stacking faults. Po-tentially more deleterious crystal defects can resultfrom large molecular impurities present in improp-erly purified source material, or even from largemolecular fractions which result from source mol-ecule decomposition on heating during the sublima-tion process. The presence of such fractions does not

depend on the quality of vacuum during deposition,but rather are a function of conditions used duringpurification and evaporation. Nevertheless, it hasbeen shown that the crystalline quality of the depos-ited film is directly dependent on the base vacuumof the system. Tanigaki et al.66 have compared thecrystalline structure of films of GaPcCl grown at arate of 0.5 Å/min on (001)KBr under base vacuumsof 5 × 10-6 and 3 × 10-9 Torr. In Figure 2-4 we showthe electron diffraction patterns of thin films grownunder these two conditions. At the lower basevacuum (Figure 2-4a), the pattern shows streakswhich are characteristic of some azimuthal disorderof the film, whereas at higher vacuums (Figure 2-4b),the more sharply defined diffraction spots indicate acorrespondingly higher degree of orientational order.In a series of experiments in our own laboratory,

PTCDA was deposited on highly oriented pyroliticgraphite (HOPG) in a conventional, bell jar metaldeposition system.1 This turbo-pumped system hada base pressure of 10-7 Torr, and the crystallinepowder source materials were loaded into baffled, Moboats prior to deposition. While it was found thathigh-resolution STM images of monolayer films couldalways be obtained from films deposited using theUHV-OMBD system, no clear images could be ob-tained for similar films deposited in the bell jarsystem. Instead, thin films deposited in the lowervacuum bell jar system appeared to clump intoislands several monolayers in thickness. While weare not certain why this difference in film order isobserved, we speculate that it is due to impuritiesincorporated during growth from the lower vacuumapparatus, along with the presence of moisture in thesource material. Given that some of the materials

Figure 2-3. (a) Steady-state mobility vs deposition rateof a 1000 Å thick film of PTCDA grown on a 0.5 W cm p-Sisubstrate at a substrate temperature of 295 K (from ref65). (b) Hole mobility in the PTCDA layer as a function ofPTCDA film deposition temperature. The mobilities areobtained from fits to current-voltage data for films ofdifferent thicknesses grown on p-Si substrates (from ref1).

Figure 2-4. Electron diffraction patterns of GaPcCl filmsgrown under (a) 5 × 10-6 Torr and (b) 3 × 10-9 Torr (fromref 66). Note the streaks which are evident in the lowvacuum sample a as opposed to the spots indicating ahigher degree of orientation in the high vacuum sample b.

1798 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 7: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

used are hygroscopic, water molecules may serve ascrystal nucleation sites, leading to island growth.Maintaining the source material at elevated temper-atures in UHV for long periods, as is done in theOMBD chamber, tends to eliminate moisture, result-ing in higher purity source material and hence higherquality monolayer films. This vacuum storage leadsto thorough drying and outgassing of the source,possibly achieving material purity over the long termfar exceeding that which is at first obtained in thethermal gradient purification process. This purifica-tion and maintenance of high material cleanlinesscannot be achieved in conventional high-vacuumsystems, where a bell jar must be opened betweenruns, thereby repeatedly exposing the source mate-rial to atmosphere.Finally, we note that there is considerable variation

as to the type of pumping used to achieve the UHVenvironment common to OMBD systems. Due to thevolatility of some organic materials, high throughputpumps are generally employed. These include dif-fusion pumps, turbomolecular pumps, and cry-opumps. In our own work, we use a high capacitycryopump during growth, with vacuum betweengrowths maintained with an ion pump. The chamberitself is lined with shrouds which are filled withliquid nitrogen to extract more impurities duringgrowth. This combination leads to a base pressureof 5 × 10-11 Torr, even after numerous depositionshave resulted in a heavy accumulation of organicmaterials on the chamber walls. In our less demand-ing work in organic thin-film devices, we employ asingle turbomolecular pump. In general, oil-contain-ing pumps such as diffusion pumps can backstream(particularly if the liquid nitrogen trap is not keptcold at all times), slowly contaminating the UHVenvironment. Furthermore, many organic materials(e.g., tris(8-hydroxyquinoline)aluminum, or Alq3) com-monly used in organic electroluminescent devices)can react with diffusion pump oil, drastically reduc-ing the pumping speed after several depositions.Nevertheless, many laboratories worldwide employdiffusion pumps as their primary means for achievingUHV environments due to their low cost and relativeease of maintenance.

3.0. Structure of Organic Thin Films Grown byOMBD

3.1. Definition of Growth ModesThe key difference between the growth of conven-

tional semiconductors and organic thin films is therelaxation of the requirement for lattice matchingwhich, in the former materials systems, significantlylimits the combination of materials which can begrown without inducing a high density of latticedefects. For example, due to the strong interatomicbonding characteristic of most semiconductor materi-als, a close lattice match (typically with strains of ∆a/aS < 10-3, where aS is the substrate lattice constant,∆a ) |aF - aS|, and aF is the lattice constant of thethin film) is required when growing films thickerthan some critical value (dc) to avoid the creation ofa very high density of misfit dislocations.67 Whilesuch mismatched epitaxy can occasionally be useful

for device applications,68 typically the carrier life-times and leakage currents which result from thehigh defect densities are unacceptable, thereby se-verely restricting the range of materials accessibleto the device engineer. On the other hand, byreducing the cohesive force of the adlayer to thesubstrate, it has been found that crystalline order canbe achieved even though the two contacting materialsare highly strained, with ∆a/a exceeding severalpercent in some cases. The materials which fall intothis class are typically bonded by van der Waals(vdW) forces, and are known as “layered” or vdWsolids.69 Perhaps the largest single class of vdWsolids are OMCs such as anthracene, perylene, thephthalocyanines and their derivatives.While the properties of OMCs have been studied

for over 50 years, it has only recently been observedthat excellent structural ordering can be achieved bydeposition on a wide range of substrates, sometimeswithout regard for the degree of strain between thefilm and the substrate.2,23,37 There has been consid-erable experimental and theoretical investigation ofOMBD growth resulting in ordered thin films, al-though currently there is no detailed understandingof the range of materials (both film and substrate),and conditions under which long-range structuralordering can be achieved. Indeed, only a looseunderstanding of the nature of the substrate/ad-sorbed layer interaction in such vdW-bonded systemshas led researchers to term ordered growth of OMCsvariously as “epitaxy”,2 “quasiepitaxy”,24 and “van derWaals epitaxy”.69

Epitaxy, as in the case of inorganic materials,refers to systems where there is a one-to-one com-mensurate relationship between the molecular posi-tions in the deposited layer and the substrate. ForOMCs, there are two types of epitaxy: conventionalepitaxy where the molecules are chemisorbed ontothe substrate surface46 and van der Waals epitaxy(vdWE) where physisorption involving only vdWbonding dominates.70 Slight mismatches between thesubstrate and film lattices results in “strained vdWE”,although for highly strained growth, epitaxial struc-tures tend to relax at a critical thickness (dc), therebygenerating defects. Epitaxy tends to occur as a resultof equilibrium growth conditions, and depending onthe relative strengths of the adsorbate-adsorbate,and adsorbate-substrate interactions, the layer growsvia one of three modes71 illustrated in Figure 3-1:layer-plus-islands (Stranski-Krastanov), layer-by-layer (Frank-van der Merwe), and island growth(Volmer-Weber). Due to the requirement for com-mensurability, it is difficult to find OMC/substratecombinations leading to unstrained vdWE, in whichcase Stranski-Krastanov growth tends to be themost frequently observed mode.In contrast, quasiepitaxial (QE) structures are

those in which the substrate and film are incom-mensurate over any meaningful lattice length scale.Nevertheless, there can be a well-defined orienta-tional relationship between the film and substratelattices resulting in azimuthal order, and in somecases residual stress develops in the film as itconforms to the substrate lattice (strained QE). Themost interesting feature of strained QE films is that

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1799

Page 8: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

they can be distorted from their bulk lattice struc-ture, although there is no significant relaxation ofthis distortion with film thickness (i.e., dc w ∞).Instead, it has been proposed that strain relaxationin QE films results from periodic variations in thedegrees of freedom internal to the adsorbate mol-ecules and their unit cells.72 This property of vdW-bonded OMCs is a direct result of their very smallelastic constants. In contrast to vdWE, QE is pri-marily achieved under nonequilibrium growth condi-tions, and since strain relief can occur withoutinducing disorder, Frank-van der Merwe as well asStranski-Krastonov growth modes have been ob-served.52,73,74We note that QE is related to two phenomena

previously reported in vdW-bonded solids: “orienta-tional epitaxy” of inert gas atoms physisorbed ontothe surface of graphite75 and the anchoring of liquidcrystal molecules onto glass or other prepared sub-strate surfaces.76,77 Indeed, the first report of QEordering also involved OMCs deposited on glasssubstrates,65 suggesting a clear link between thesephenomena.Figure 3-2 is a diagram placing these several

growth modes in relative context.78 The vertical axisshows the logarithm of the bond energy to thesubstrate, where vdW bonds are typically in therange of 1-10 meV/atom79 (although the bindingenergy/molecule can exceed 1 eV), whereas for cova-

lent or ionic bonding, the energies range from 100meV to ∼5eV/atom.80 Unstrained, epitaxial growthcorresponds to ∆a/a ) 0. The curved line schemati-cally indicates the strain for a constant critical layerthickness:67 as the bond energy increases, strain mustdecrease if dc is to be maintained at a constant value.At higher strains, relaxed polycrystalline or amor-phous growth occurs. While different regions of thisenergy/strain plot are occupied by strained vdWE andQE, there are also areas of significant overlap. Thatis, for a particular strain energy, strained vdWE orQE structures can be attained, depending on thegrowth conditions employed.In this section, we will discuss examples of OMC

thin film structures grown into both strained vdWEand QE structures. There are, to our knowledge, nounambiguous examples of completely unstrained,heteroepitaxial vdWE of OMCs, primarily due to alack of suitable substrates whose lattice constantsand space group are matched to that of the adsorbedlayer. To distinguish between these growth modes,very high resolution materials characterization isneeded. Unfortunately, given the inherent fragilityof many of the materials used, and the unusualstructural nature of the resulting films, very fewhigh-resolution tools have been developed, and fewfilms have been completely characterized. To date,the most thoroughly studied system is that based onthe the planar organic molecule, PTCDA, which hasbeen the focus of considerable study due to itsregular, planar stacking properties on a variety ofsubstrates.1,37,46 This is a particularly interestingsystem on which to focus our attention since bothepitaxial and QE films of PTCDA have been grownunder a range of growth conditions. Several phtha-locyanine and porphyrin-based systems have alsoreceived much attention.33,81 These and other ex-ample materials will be used to illustrate the un-precedented range of growth modes and structureswhich are accessible via the combination of OMBDand organic molecular solid-state systems.

3.2. Epitaxy and van der Waals EpitaxyThe most extensive studies of epitaxial growth of

OMC films have concentrated on the phthalocyanines(Pc) layered on ionic substrates such as the alkalihalides,82-89 metals,90 and on passivated surfaces53,81,91of Si(111) and GaAs(111). Other systems consistingof C60 deposited on various substrates such as mica92,93and semiconductor substrates94 have also been stud-ied in some detail. Typical conditions leading toepitaxial growth of planar molecules include lowgrowth rates (<0.001 to 0.01 ML/s) and high sub-strate temperatures (∼100-200 °C, depending on thesublimation temperature of the organic molecule).These conditions result in the growth of equilibrium

Figure 3-1. Various modes of epitaxial growth: (a) layer-by-layer (Frank-van der Merwe), (b) layer-plus-island (Stranski-Krastonov), and (c) island (Volmer-Weber). The horizontal lines on the epitaxial layers schematically represents individualatomic or molecular layers.

Figure 3-2. Schematic diagram of modes of epitaxy.Epitaxial growth occurs for commensurate lattices whileQE growth occurs for vdW bonded incommensurate thinfilms. The left-hand axis corresponds to unstrained growth.The curved line represents the boundary between orderedand disordered growth assuming a particular value ofcritical thickness, dc. The lower the binding energy is, thelarger the strain allowed to achieve unrelaxed growth of agiven thickness, d < dc (from ref 78).

1800 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 9: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

thin film structures, where the incident moleculeshave sufficient thermal energy after deposition toarrange themselves into their minimum energy con-figuration.Epitaxial molecular thin films are typically grown

into structures which are determined by the sub-strate lattice rather than by the bulk structure of theorganic molecular solid itself. Since the substratesused are typically inorganic crystals whose structureis considerably different than the OMC, neither thelattice constant nor the lattice symmetry of these twocontacting layers are matched. This has twoconsequences: due to the flexible nature of the vdWforce responsible for the cohesion of the organiclayers, the structure of the thin film can be substan-tially distorted from bulk. This severe lattice distor-tion, however, results in a large stress within the

film, thereby creating disorder after the growth ofonly a few monolayers.46,83 In effect, the mismatchedfilm relaxes to its bulk structure within 1-5 ML,resulting in a rough columnar surface morphology(corresponding to Stranski-Krastanov growth71). Thefilm relaxation prevents the growth of molecularlyflat films with thicknesses beyond only a few mono-layers.An example of organic thin-film epitaxy is the

OMBD growth of PbPc on KBr(001) and NaCl(001)as studied using in situ reflection high-energy elec-tron diffraction83 (RHEED). For these experiments(summarized in Table 1), the growth rate was from0.003 to 0.008 ML/s with the substrate maintainedat room temperature. Figure 3-3a shows the molec-ular structure of PbPc, where the Pb atom forms anout-of-plane pyramid with the surrounding ligands.

Table 1. Surface Unit Mesh Lattice Constants for Pc’s on Different Substratesa (From Refs 83 and 97)

molecule bulk (nm) KBr (nm) KCl (nm) KI (nm) NaCl (nm) MoS2 (nm) Se-GaAs(111)B (nm) H-Si(111) (nm)

VOPc 1.40 1.41 1.26 1.37 1.37 1.38AlPcCl 1.48 1.41 1.50 1.43 1.44 1.38PbPc 1.27 1.40 1.41 1.26a All distances in nanometers.

Figure 3-3. (a) Molecular structural diagram of PbPc. Proposed epitaxial orientations of PbPc on (b) NaCl and (c) KBr.For PbPc on KBr, two different arrangements are proposed: commensurate type A (left), and incommensurate type B′(right) (from ref 83).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1801

Page 10: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

The PbPc bulk monoclinic lattice forms a com-mensurate square lattice when deposited onto severalalkali halide crystal surfaces. For example, Figure3-3, parts b and c, shows the first ML structure ofPbPc on NaCl and KBr, respectively, as inferred fromthe RHEED patterns obtained immediately aftergrowth. To the accuracy achieved using RHEED((5%), the lattice constants are considerably differentin these two cases. For example, for PbPc on NaCl,the lattice constant is aF ) 1.26 nm (comparable tothe bulk lattice parameter of 1.27 nm), whereas forKBr, this distance is aF ) 1.40 nm, or a difference of11%. In the case of growth on KCl(001), com-mensurability is achieved by the PbPc forming twoequivalentx10×x10-R(18.4° (type B) surface mesheswith a considerably expanded unit cell of aF ) 1.41nm. Here, we use the conventional notation to definemolecular surface reconstructions. Thus, (m×n)denotes a surface unit mesh which is m×n timeslarger than the underlying lattice. If rotated fromthe underlying lattice by θ°, then the notation is(m×n)Rθ°. If the surface mesh is centered on (ratherthan commensurate with) the bulk lattice, the nota-tion is c(m×n).On KBr, two different lattices are observed: the

3×3 (type A) commensurate lattice mesh with aF )1.40 nm coexists with a much smaller proportion ofan incommensurate (i.e., quasi-epitaxial) type B′lattice where the PbPc unit cell is parallel to the [210]substrate axis. Many organic materials, such as thePc’s, have more than a single structure. Hence, whileinvestigations of PbPc do not identify more than asingle commensurate, but highly strained isomorph,we cannot entirely rule out the possibility that a lessstrained (i.e., more commensurate) form of PbPc hasactually been achieved. Nevertheless, lacking theability for growth of smooth films more than a fewmonolayers thick provides evidence for highly strainedstructures. To accommodate the significantly ex-panded PbPc structure when deposited onto KBr, itwas suggested83 that the molecules rotate such thatthe phenyl group of one molecule fits into the “hollow”of an adjacent molecule to minimize the intralayervdW potential. In all cases discussed, steric conges-tion between neighboring molecules causes them tominimize their vdW binding energy by rotating aboutan axis normal to the molecular plane, with thedegree of rotation determined by the size of thesurface mesh required to achieve commensurability.Utilizing the additional degrees of freedom within theunit cell, the molecules remain parallel to the sub-strate.The molecule-substrate bonding is primarily due

to electrostatic interactions between the metal atomwith the anion in the substrate lattice. The impor-tance of electrostatic interactions in determiningstructure is inferred from extended Huckel modelcalculations, which suggest that a charge of from 0.4to 0.6 e is localized on the Pb atom. Hence, theresulting epitaxial structure of Pc thin films isdetermined by two factors: strong electrostatic at-traction between the adlayer and the substrate, andvdW forces within the layer itself which are mini-mized by the molecular rotational adjustments notedabove. The importance of electrostatic attraction to

the ionic substrate is apparent when we observe thatthe lattice constants of the surface unit mesh ofdifferent organics on the same substrate (e.g., KCl)are the same (Table 1), independent of the equilib-rium (bulk) constant of the thin film. The total strainenergy increases with film thickness until the epi-taxial layer relaxes, resulting in a highly disorderedstructure after only a few monolayers (typically < 5ML, depending on the degree of strain).Note that the electrostatically bonded thin films

are not consistent with the definition of vdWE. Insome cases, it might be possible to grow one or twostrained “wetting” layers which are bonded electro-statically, followed by the vdWE or QE growth ofsubsequent layers to form an ordered thin film,resulting in Stranski-Krastonov layer-plus-islandmorphology. In one approach to reducing substrate-thin film interactions, several Pc molecular systemshave also been grown on H-terminated Si(111) andSe-terminated GaAs(111) surfaces using conditionssimilar to those employed for the growth of Pc’s onthe alkali halides.53,91 Atomically flat, regularlyterminated monohydride (-SiH) Si(111) surfaces canbe prepared by immersing the clean Si substrate inHF solution immediately prior to growth. Growthof VOPc on such a surface resulted in a nearlyequilibrium (unstrained) square lattice with aF ) 1.38nm. Long range structural order could then bemaintained for thicknesses as large as 20 ML,indicating low stress in the first few anchoring layers.In contrast, a lattice constant of aF ) 1.38 nm is toosmall to easily fit the larger AlPcCl molecule. Theresulting stress in these latter films generates poorlyordered layers of even a few ML thickness.The perylene-based molecule, PTCDA, has also

been found to grow epitaxially on alkali halides, withgrowth extending to very thick films.46 To accom-modate the strain, however, highly textured colum-nar structures result. This Stranski-Krastanovgrowth is stable even for thick films since the domainsize is small. Once the growth is nucleated at thesubstrate-film interface, the lattice structure almostimmediately relaxes into its bulk crystal habit,resulting in numerous small domains which growwithout significant strain since the contact surfacearea of the domain with the substrate is also small.Epitaxy of PTCDA on NaCl(001) and KCl(001)

substrates has been studied extensively by Mobus etal.46 where equilibrium growth was achieved bymaintaining the substrate at 200 °C at a growth rateof 0.08 Å/s (corresponding to ∼1.5 ML/min). Underthese conditions, the slowly arriving molecules havesufficient thermal energy to diffuse along the sub-strate surface until they find a minimum (equilibri-um) energy site. This results in highly columnargrowth, with voids arising between neighboringislands. Within the islands, the molecules formregular stacks, with the PTCDA (102) plane lyingparallel to the substrate surface.Using transmission electron microscope diffraction,

it was found that PTCDA grows into two, coexistingR and â polymorphs. The appearance of these twomodifications of the PTCDA unit cell (whose latticeconstants are provided in Table 2), shown in Figure3-4, suggest that the strong substrate-organic film

1802 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 11: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

interaction (when compared to the intralayer interac-tion within the OMC itself) in epitaxially grownorganic films plays a determining role in the resultingPTCDA structure, similar to that observed for thePc’s. This interaction also determines the azimuthalorientation of the crystalline domains, as the adlayerminimizes energy by finding a commensurate orien-tation with the substrate. Several orientations of themolecules within the islands relative to the substratelattice have been observed, and these are shown inFigure 3-4 for R- and â-PTCDA on NaCl and KCl.For example, on NaCl, the [0,1] unit mesh axis of

R-PTCDA is rotated by θ ) 4° with respect to the[110] NaCl axis, whereas for â-PTCDA, θ ) 7° or 31°.For all of these arrangements, the four carbonyloxygen atoms align to a Na+ ion in a commensuratesuperstructure which can repeat over distances ofseveral PTCDA unit cells. Due to the high latticesymmetry, there are eight symmetry-equivalent ori-entations for each of the two forms of PTCDA onNaCl, and all eight are observed after a single growthrun. The situation for PTCDA on KCl is similar toNaCl, except that the preferred orientations for bothR and â forms is θ ) 24°, which provides a nearlyperfect epitaxial alignment of two opposite O atomswith the K+ ions on the substrate. This preferredalignment to the substrate cations suggests thatelectrostatic interactions play a role for PTCDAgrowth on alkali halides similar to that found for thePc’s.Epitaxy of PTCDA on metals has also been inves-

tigated by Seidel and co-workers.95 In that work,monolayers of PTCDA were deposited onto pre-cleaned (by sputtering with 500eV Ar+ ions and

Table 2. Unit Cell Parameters of PTCDA and NTCDA

parameter R-PTCDAa â-PTCDAb NTCDAa

space group P21/c(C2h5) P21/c(C2h

5) P21/c(C2h5)

a (Å) 3.72 3.78 7.89b (Å) 11.96 19.30 5.33c (Å) 17.34 10.77 12.74â 98.8° 83.6° 109.04°Z 2 2 2a From ref 99. b From ref 46.

Figure 3-4. R and â isomorphs of PTCDA and their proposed orientations when grown on NaCl and KCl. In all cases,the PTCDA (102) lies parallel to the substrate surface. These orientations are regarded as tentative, since no precisemicroscopic registry has been determined in all cases (from ref 46).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1803

Page 12: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

heating to 1000 K) Ag(110) surfaces. Under theconditions used, epitaxy (rather than coincident orquasi-epitaxial growth) was inferred, although as willbe shown in section 3.3.2.3, the growth mode ofPTCDA achieved on metal substrates depends sen-sitively on the conditions used in the process.For the case of PTCDA/Ag(110), growth occurred

at a background pressure of <4 × 10-10 Torr and ata growth rate of∼7× 10-3 ML/s. The same structurewas obtained for substrate temperatures rangingbetween 200 and 380 K during growth. The resultingML films were examined by both LEED and STM todetermine the unit cell structure and size. LEEDdata were taken at near normal incidence using abeam energy of 16 eV assuring no damage wasincurred by the film during measurement. To withinthe accuracies of the measurement techniques used(which typically are ∼5-10%), it was found that acommensurate superstructure was aligned along theAg (100) lattice direction, with the following relation-ships between the PTCDA unit vectors (a*, b*) andthose of Ag (g1, g2):

Lack of Moire patterns in the STM images providesadditional confirmation that the PTCDA lattice iscommensurate with the underlying Ag(110) lattice.96Interestingly, the structure of the first ML of

PTCDA is strongly distorted from its bulk structure,similar to the case observed for several Pc’s grownepitaxially on alkali halide substrates discussedabove. This is apparent in the STM images95 ofPTCDA ML’s grown on Ag(110) terraces shown inFigure 3-5a. For example, as opposed to the normalherringbone structure of bulk PTCDA with twomolecules per surface cell (with dimensions a ) 11.96Å and b ) 17.34 Å and â ) 98.8°, given by the R formin Table 2), epitaxy on Ag(110) results in a surfaceunit cell where the molecules are in a nearly squarelattice with Z ) 1, a* ) 11.8 ( 0.5 Å, b* ) 12.5 (0.5 Å and â ) 83.2°. These results were consistentwhether obtained via LEED or STM. Occasionallya defect in the lattice was observed, as indicated bythe arrow in Figure 3-5a, where one monolayerappears to have “reverted” into its bulk, herringbonestructure. This is indicative of a competition betweenminimizing energy within a layer as opposed tominimizing energy between the layer and the sub-strate. The critical balance between these factorsgoverns the ultimate growth mode of organic thinfilms as it does in the case of inorganic materials,and will be discussed in greater detail in the followingsections.This extreme distortion of the unit cell, which was

also observed in â-PTCDA discussed previously, isdue to the strong covalent interactions between thecarbon and silver atoms in the monolayer andsubstrate, respectively. The strength of this interac-tion was confirmed by thermal desorption of multiplePTCDA layers previously deposited on Ag. It wasfound that while the upper layers could be desorbedwithout decomposition, desorption of the final layerresulted in molecular fragmentation due to strongsubstrate-molecule adhesion.

Minimum energy calculations were also used toprovide a quantitative understanding of the observedstructures.95 For these calculations, the vdW energy,nonbinding electrostatic energy between atoms in themolecule and substrate, bond stretch and angulardistortion energies, and dihedral angular torsionwere summed to obtain the total adlayer energy andthen minimized using standard methods. Thesecalculations confirmed the distorted structure shownin Figure 3-5b, with a lattice of a* ) b* ) 11.92 Åand â ) 86.67°, which is reasonably close to themeasured values.To summarize the results given in the examples

of the Pc’s and PTCDA, epitaxy is achieved undergrowth conditions which allow for an equilibrium,lowest energy structure to nucleate. Typically, thisoccurs for elevated substrate temperatures, lowgrowth rates, and strong substrate-adlayer interac-tions (e.g., ionic bonding for PTCDA on alkali halidesubstrates). In the case of Pc or PTCDA growth onalkali halides, or of PTCDA on Ag(110), it can beshown that the unit cell of the adlayer can besignificantly distorted from its bulk dimensions in

a* ) 3g1 + 2g2

b* ) 3g1 - 2g2

Figure 3-5. (a) Scanning tunneling microscope image ofseveral terraces of PTCDA on a Ag(110) surface. Arrowindicates row where the molecular orientation is reversedfrom other rows on that same substrate terrace. (b)Proposed comensurate superstructure of a ML of PTCDAon Ag(110), where the Ag(110) surface mesh is indicatedby the dots, with lattice paramters g1 ) 2.089 Å and g2 )4.08 Å (from ref 95).

1804 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 13: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

order to achieve a commensurate epitaxial structure.While this allows for the growth of a highly strainedinitial ML, thicker films tend to relax, therebyinducing a high density of defects along with a roughsurface. In the case of PTCDA with its very strongintramolecular binding, the molecules azimuthallyorient with respect to the substrate to achieve acommensurate, or nearly commensurate structure.The experimental resolution of these experi-ments46,66,82,83,95,97 is not sufficient, however, to de-termine the degree of strain resulting from thisovergrowth. Nevertheless, the epitaxial PTCDA lay-ers exhibit highly textured, island growth resultingfrom different domains of different polymorphs of thePTCDA structure nucleating in spatially distributedregions of the substrate. This mode of strain relax-ation appears to be a universally observed charac-teristic of strained-vdWE organic thin film growth.Finally, some recent attempts have also been made

to find substrates whose lattice parameters closelymatch those of the overlayer in order to achievenearly unstrained vdWE. To our knowledge, the bestexample of such growth98 is that of the nonlinearchromophore, 4′-nitrobenzylidene-3-acetamino-4-meth-oxyanaline (MNBA) on the (010) plane of nearlylattice-matched single crystals of the organic salt:ethylenediammonium terephthalate (EDT). The lat-tice constants of the two bulk crystals in the (010)plane are a ) 8.3814(9) Å and c ) 7.4871(5) Å withâ ) 115.336(5)° for MNBA, compared with a )8.381(8) Å and c ) 7.471(4) Å with â ) 115.32(5)°for EDT. Thus, at room temperature, the maximumstrain is ∆a/a < 0.3%.Single crystals of EDT were prepared by slow

evaporation or by controlled supercooling out ofsolution. This yields (010) surfaces (determined byX-ray diffraction) in the millimeter range. Prior togrowth, the surfaces were cleaned using thermaldesorption in vacuum by ramping the temperatureup to 120 °C with a dwell time of 30 min, at whichtime the heteroepitaxial layer of MNBA was grown.Optimum growth of MNBA was obtained at a back-ground pressure of 10-9 Torr, a substrate tempera-ture of Tsub ) 80 °C, and a rate of 0.1-0.2 Å/s. Theseconditions resulted in island growth morphology,with asymmetric islands typically 1 × 10 µm2 ori-ented normal to the [001] direction of the EDTsubstrate. The islands appear to fill less than 50%of the substrate surface area. It was also noted98 thatthe range of temperatures over which this degree oforder could be obtained was quite narrow: below Tsub) 75°, the island orientation became random, whereasat Tsub > 85 °C, the desorption rate became so highsuch that MNBA no longer was deposited.The island orientation is clearly apparent in optical

micrographs, and was confirmed by the strong de-pendence of the second harmonic intensity generatedby MNBA on the fundamental optical pump beam(at a wavelength of λ ) 1.064 µm) polarization angle.It was found that the maximum second harmonicsignal was obtained for polarizations parallel to thesubstrate [100] direction, while the signal was re-duced by a factor of 90 for perpendicular polariza-tions. From this result, it was concluded that the[100] directions of EDT and MNBA were coincident.

3.3. Quasi-EpitaxyThe most thoroughly studied quasi-epitaxial sys-

tem is PTCDA grown on a variety of substrates,1,37,99including highly oriented pyrolytic graphite(HOPG),43,58,70,100 Au(111),73 Se-terminated GaAs(100),55,101 glass,37 and polymers.102 Additional ex-amples of this growth mode include several ad-ditional examples of neutral organic molecules onamorphous quartz,103 and HOPG,96,104 C60 and C70 onGaSe and MoS2,105 and possibly C60 on Au(111).106 Incontrast to epitaxy, QE growth is a kineticallycontrolled, nonequilibrium process, resulting in struc-tures which can be significantly distorted from thebulk. The conditions leading to kinetically controlledgrowth are high deposition rates and low substratetemperatures. An ordered film structure is obtaineddue to dominant intralayer molecule-molecule inter-actions (as compared to epitaxy, where interlayerforces dominate).

3.3.1. Theory of QEIn this section, a theoretical framework and ex-

perimental basis for QE growth of archetype vdW-bonded molecular thin films is presented. The modelwhich has been developed by Zhang and Forrest72 iscalculationally simple in that it replaces the largenumber of pairwise atom-atom potentials whichmust be summed between adjacent molecules by asingle, approximate, ellipsoidally symmetric molecule-molecule potential. We note, however, that theapproach is inherently limited since it only predictsequilibrium structural configurations without provid-ing information as to the growth conditions requiredto achieve a given structure. To date, these condi-tions have been experimentally discovered via aprocess of trial and error. The model has thus farbeen applied to structures consisting of the planarmolecules PTCDA and 3,4,7,8-naphthalenetetracar-boxylic dianhydride (NTCDA), with results in goodagreement with observation.While models have also been previously developed

to understand epitaxial growth using approximationssimilar to those employed here (e.g., using the rigidlattice approximation107,108), epitaxial systems areinherently different from QE systems. The primarydifference lies in the incommensurability of QE layerswith the substrate. Whereas one can model epitaxialgrowth by a harmonic potential with a period equalto the atomic spacing of the substrate,107 incom-mensurate lattices cannot be treated as such sincethe potential between overlayer and substrate isanharmonic. Hence, the analytical solutions whichare attained for epitaxial systems must be replacedby computationally intensive methods.Extending models of purely van der Waals-bonded

molecules to systems where long-range (but weak)intra- and interlayer Coulomb binding plays a rolecan significantly complicate the problem, and has notbeen treated in detail. However, the calculationalmethods employed for vdW systems can, in principle,be extended to include these and other bonding forces(e.g., hydrogen bonding, high order multipoles, ionicand covalent bonds, etc.).Interestingly, inclusion of Coulomb forces between

the molecule and the substrate only complicates the

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1805

Page 14: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

calculation when these forces do not dominate. Whensuch interactions are strongly dominant, as in thecase of VOPc on alkali halide (and hence ionic)substrates, conventional epitaxy results (section 3.2).In this case, intralayer interactions between mol-ecules, which are largely responsible for determiningthe bulk structure of OMCs, are not significant, andthe equilibrium configuration of the first monolayeris accurately determined by the single molecule/substrate binding energy.109 In effect, this is thedefining difference between epitaxy and QE. Asnoted above, this strong substrate-molecule bindingcharacteristic of epitaxial growth results in thegeneration of a high density of defects, leading toamorphous or highly polycrystalline films thickerthan few monolayers.The primary requirement for QE is that there

exists a range over which a surface molecule can betranslated relative to the substrate without a sig-nificant change in energy. If the potential betweenmolecules within a layer is φintra, and between mol-ecules in different layers is φinter, then this conditionis related to the relative magnitudes of the inter- andintralayer compressibilities (or elasticities) via

where φ′′ is the second derivative of φ relative to theseveral thin-film spatial degrees of freedom. In this“static” approximation, the condition for QE is rela-tively independent of the absolute magnitudes of thecrystal binding energies, φintra or φinter. As in manycases involving large planar molecules, φinter > φintrawhile φinter′′ << φintra′′. This differs from atomic vdWsystems where there is no strict adherence to theseconditions.110 This point is noteworthy since it il-lustrates an inherent difference between the previ-ously studied “orientational epitaxy” of inert gasatoms vdW-bonded to metal and graphite surfaces,75and QE of spatially extended molecules bound bysimilar forces on a substrate with considerably higherspatial frequency (due to its smaller lattice constant)than the adsorbed molecular layer.3.3.1.1. Based on Energy Minimization. The con-

dition in eq 3.1 has been examined by calculating thevdW bond energy between 2D interfaces of PTCDAand HOPG using the atom-atom potential method.79Here, the total bond potential is given by Φ ) ∑φij,where φij is the potential between the ith and jthatoms in two molecules, or in a molecule and thesubstrate. Now, φij can be approximated for vdWsolids using the Buckingham potential:

The radial distance between atoms i and j is rij, andR, â, and γ are vdW constants for each pair of atoms.The minimum energy corresponding to the equilib-rium crystal configuration is obtained when Φ′ ) 0and Φ′′ > 0.There are several assumptions made when apply-

ing the atom-atom potential procedure:79 (i) Themodel is static. That is, the calculation applies toequilibrium at T ) 0 K, and hence it cannot predictdynamic growth processes, nor does it considervibrational contributions at T > 0. Typically, these

latter effects are small compared with Φ. (ii) ThevdW forces are isotropic, and are not significantlyperturbed by the molecular structure. (iii) Themolecules are rigid (i.e., there are no internal mo-lecular degrees of freedom. (iv) There is no signifi-cant contribution to the intermolecular energy arisingfrom Coulombic forces, higher order multipoles, etc.Previous calculations for planar, nonpolar molecules79such as anthracene and coronene have shown thatthese assumptions are generally valid.The condition for QE (i.e., φintra′′ >> φinter′′) can be

understood from the plot of elastic constants betweenindividual PTCDA molecules shown in Figure 3-6,where the compressibility within a layer (propor-tional to φintra′′) is clearly larger than the elastic shear(proportional to φinter′′) near the equilibrium molec-ular separation of ∆x ) 0. In this case, the overlayeris only slightly distorted when deposited onto asubstrate, allowing for some adjustment of the posi-tion of the relative layer orientation without inducinga large strain energy. In effect, the potential surfaceof the molecule-substrate interaction (shown in theinset of Figure 3-6) has a relatively broad minimum(leading to a small φinter′′), allowing for the requiredorientational adjustment between lattices without alarge expense of strain energy. This broad energyminimum is a direct result of the short range vdWbinding energy, and the spatial extent of planarmolecules such as PTCDA and the Pc’s. That is, themore extended molecules result in a higher interfa-cial compressibility (i.e., lower elasticity, φinter′′).Such molecules have a broader range of energy-equivalent positions, and therefore can result in thegrowth (under appropriate conditions) of an incom-mensurate overlayer with only a small interfacialstrain energy, as required for ordered QE growth.These assumptions lead to a picture of a QE thin

film as a rigid overlayer, relatively undistorted whenplaced in contact with the incommensurate substrate.This rigid overlayer approximation greatly simplifiescalculations of the full QE structure, as will be shownbelow. Since the overlayer has a very low spatialperiodicity as compared with the substrate, strainmay be relieved through modifications of the internal

φintra′′ . φinter′′ (3.1)

φij ) -Rij/rij6 + âij exp(-γijrij) (3.2)

Figure 3-6. Calculated interlayer (φinter′′) and intralayer(φintra′′) compressibilities as a function of displacement fromequilibrium (∆x) of a PTCDA crystal. The condition forQE: φintra′′ . φinter′′ at ∆x ) 0 is clearly achieved for thismaterial system (from ref 78).

1806 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 15: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

degrees of freedom within the larger, adsorbed mo-lecular lattice. However, as will also be shown in thefollowing section, precision structural measurementssuggest that QE films can have significant strainwhen grown on substrates where φinter′′ ≈ φintra′′. Theresulting elastic distortion of the film as comparedto the bulk lattice structure does not appear to relaxeven in the thickest films studied.111 This suggeststhat the details of QE structures are considerablymore complex than predicted by the model discussedhere, although the general features of QE structuresare indeed predicted by the somewhat simplifiedassumptions used to make the calculations bothintuitively useful and calculationally tractable.The accuracy of the method has been tested by

calculating the bulk crystal structure of PTCDA, andthen comparing these results with existing crystal-lographic data. Note that calculation of the full bulkstructure is a complex 3D problem involving numer-ous degrees of freedom between the several moleculesin the cells. Hence, only limited aspects of the 3Dstructure are calculated to test the model. Perspec-tive views of the bulk PTCDA and NTCDA unit cellsare provided in Figure 3-7, parts a and b, respec-tively,63 with details of the structures compiled inTable 2.The energy of two organic molecules stacked one

above the other was calculated as a function ofintermolecular distance along the stacking axis, using

previously published values for the atomic vdWcoefficients,112,113 with the results plotted in Figure3-8. Bonding energies for PTCDA-PTCDA, NTCDA-NTCDA, and PTCDA-NTCDA dimers are all plottedin the figure. It is apparent that the equilibriumdistance (or “vdW radius”) in the stacking direction(assuming that the molecular planes of adjacent

Figure 3-7. Perspective views of the unit cells of (a) PTCDA and (b) NTCDA.

Figure 3-8. The van der Waals potential as a function ofinterplanar stacking distance of a PTCDA dimer, a NTCDAdimer, and a PTCDA/NTCDA stack. For each curve, thetwo molecules are centered with respect to each other, withtheir molecular planes parallel (from ref 99).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1807

Page 16: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

molecules are parallel, which is not the case for theNTCDA-NTCDA bulk crystal) are generally in therange from 3.20 to 3.26 Å, corresponding to theexperimentally determined value of 3.21 Å for PTC-DA. Furthermore, the binding energy of the PTCDAdimer is approximately -1.3 eV, corresponding to 24kcal/mol. This value is consistent with the sublima-tion energies114 of many aromatic molecules similarto PTCDA and NTCDA.Once the stacking distance along the c axis is fixed,

the two molecules are translated in the xy (base)plane with respect to each other to generate theenergy surfaces for PTCDA and NTCDA dimersshown in Figure 3-9, parts a and b, respectively.There are degenerate energy minima at (1.1 Å fromthe center of the PTCDA molecules, and (1.0 Å forthe NTCDA molecules. The energy minima aredisplaced from the molecular center of mass and arelocated along the long molecular axis. The equilib-

rium molecular offset is experimentally observed inthe PTCDA bulk structure, resulting in the molecularlamella stacking along the (102) axis. That is, thereis an 11° tilt of the molecular plane within the unitcell, which has been widely observed for PTCDAdeposited on such insulating substrates as glass65 andpolymers.22,115 Interestingly, this 11° tilt is notobserved for PTCDA on conducting substrates suchas Au or GaAs, suggesting that more complex chargeexchange interactions not included in these calcula-tions may exist for these situations.The energy of a 2D surface mesh consisting of five

PTCDAmolecules placed in their characteristic base-centered rectangular structure, shown in Figure 3-10,was also calculated. For this calculation, all anglesand dimensions were varied (including the relativeangle of the central molecule with respect to theplane defined by the four corner molecules) to achievethe minimum energy configuration, with the result

Figure 3-9. The 2D dimer energy surface for (a) PTCDA and (b) NTCDA for the molecules oriented with their planesparallel (from ref 99).

1808 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 17: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

shown in the figure. The angles and dimensions ofthe calculated surface unit cell agree well with theactual surface unit as determined from RHEED andSTM (see Table 3). We note that the surface unitcell is enlarged by 25% over that obtained for a bulkcell, with the calculated minimum energy corre-sponding to xmin ) 20.0 ( 0.5 Å and ymin ) 15.7 ( 0.5Å. This accounts for (although it somewhat overes-timates) the reconstructed surface dimensions ofPTCDA observed by both STM and RHEED. Thesurface cell minimum energy is obtained for allmolecules in a coplanar configuration. By placing asecond layer at 3.21 Å above the surface, the cellminimum energy is achieved only when xmin and yminare decreased to values approaching their experi-mentally obtained bulk distances of xmin ) 17.34 Åand ymin ) 11.96 Å. Also, the central molecule of thesurface cell is rotated to an angle of ê ) 0.48 ( 0.02rad, also consistent with observation.The orientation of the adsorbed layer with respect

to the substrate lattice is defined by angle, θ, betweenthe primitive vectors in the two lattices (see Figure3-10). From this figure, θ is a well-defined quantityfor all epitaxial systems; whether they are com-mensurate, incommensurate, or coincident.104 Thevalue of θ leading to Φ′(θ) ) 0 and Φ′′(θ) > 0

corresponds to the equilibrium lattice configuration.On the basis of these arguments, and using the “rigidlattice” approximation where the layers are trans-lated with respect to each other without distortion,θ has been calculated for a monolayer of PTCDAdeposited on graphite,99 with the results shown inFigure 3-11. Here, a sufficiently large lattice (10×10PTCDA unit cells, or ∼50 000 atoms) was consideredto ensure insensitivity of the result to layer bound-aries and variations due to center of mass translationof the overlayer. An energy minimum of -4 meV wasnoted at θmin ) 0.85 rad, where the energy axis inFigure 3-11 is normalized to a single adsorbedmolecule of PTCDA (-2.0 eV). This relatively smallenergy perturbation/molecule (<0.3%) justifies theassumption that the overlayer structure is not stronglyinfluenced by the substrate, in contrast to atomicvdW systems (e.g., Ar on graphite) where the rota-tional energy is∼5% of the bond energy.116 Note thatθmin is in relatively good agreement with the experi-mental value43,58,70 of θ ) 0.84 ( 0.07 rad (see Table3) obtained by STM.Both RHEED and X-ray data52,73 indicate unit cell

dimensions that are somewhat smaller than pre-dicted by this simple theory. This discrepancy is inpart due to formation of an expanded first ML, andthen subsequent relaxation into a structure closer(but not identical) to the bulk crystal habit of PTCDA.A transition interfacial layer of dimensions differentfrom subsequent layers appears to be one mechanismfor the relief of strain in these incommensurate QEsystems; a feature noted in studies of the transitionfrom equilibrium growth (i.e., at high substratetemperature, Tsub) to nonequilibrium, or QE, growth(at low Tsub) of PTCDA on Au(111) (see section3.3.2.3).

Figure 3-10. Surface unit mesh of PTCDA on a graphitesubstrate defining angles and dimensions referred to in thetext.

Table 3. Surface Unit Cell Parametersa for PTCDA on Graphite

a (Å) b (Å) σ (rad) ê (rad) θ (rad) ref(s)

theory 20.0 ( 0.5 15.7 ( 0.5 0.86 ( 0.02 0.48 ( 0.02 0.86 ( 0.02 72,99STM 20.5 ( 1.2 14.2 ( 1.5 0.96 ( 0.03 0.61 ( 0.17 0.90 ( 0.06 43,58,70RHEED 22.4 ( 1.0 16.0 ( 1.0 52,70,73

a Dimensional parameters defined in Figure 3-10.

Figure 3-11. Binding energy of a monolayer of PTCDAon a graphite substrate as a function of relative latticeangle, θ, as defined in Figure 3-10. The vertical axis is theenergy normalized to the binding energy per PTCDAmolecule (from ref 99).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1809

Page 18: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

This calculation was also used to study molecularinterfaces consisting of NTCDA on a crystallinePTCDA substrate.99 Experimental evidence obtainedfor alternating multilayer stacks of these materi-als23,70 imply that the layers form ordered crystallineorganic multiple quantum well structures. Due tothe extended size of the molecules in both layers andto the large number of pairwise atomic interactionsbetween each molecule, the method outlined in eq 3.2is calculationally impractical. For example, to de-termine the intermolecular potential between justtwo PTCDA molecules requires 1444 pairwise inter-action potential calculations. The calculations havebeen substantially simplified by Zhang72 by replacingthe ∼103 atomic interactions between each NTCDAand PTCDA molecule with an approximate, single,elliptically symmetric molecular potential assuminga fixed intermolecular stacking distance. One suchpotential which has been shown to provide a good fitto molecules within a given layer is similar to theBuckingham potential, viz.:

where rij is now the center-center distance betweenmolecules, δ(θ) is the distance of “closest approach”of two adjacent molecules as calculated from theircore repulsion, and the vdW constants are functionsof the angle (θ) between unit cells in the overlayerand substrate. An alternative potential form whichalso provides a good fit to the atom-atom calculation,especially for molecules between planes, was ap-proximated using

where the distance in the z direction is held fixed atthe interplanar spacing of d. The angular depend-encies of the parameters such as R, â, δ, etc. weredetermined by translating one molecule over thesubstrate at a fixed z (equilibrium) distance in boththe a and b crystalline directions while calculatingthe full atom-atom potential using eq 3.2. Thisprovides values for Ra, Rb, etc., or similarly for theparameters in eq 3.4. To determine these parametersat other off-axis angles, an ellipsoidal dependencewas assumed:

This calculational technique has been shown to bereasonably accurate (to within (5% of the measuredcrystal dimensions), allowing for the semiquantita-tive determination of equilibrium film-substratestructures consisting of relatively complex mol-ecules.1 In any case, we note that the accuracy ofusing a full atomistic potential is limited as well.Commonly, the values of the various vdW coefficientsused in such pairwise calculational methods areadjusted to match the predicted to the observedcrystal structure to reduce inherent inaccuracies.Similar “iterative” procedures can also be applied tofinding the optimal values of the effective coefficients

derived from eq 3.5, although the overall accuracy ofthis procedure is found to be adequate for predictingthe observed QE systems to which it has beenapplied.Recently, it has been suggested117 that while the

ellipsoidal approximation is well adapted to thecalculation of structures involving simple, smallplanar molecules such as PTCDA, it may not beappropriate for larger molecules, or molecules whichexhibit some out-of-plane (3D) structure, such as thephthalocyanines. To generalize these approxima-tions to account for complex molecular shapes, Liuet al.117 represent the molecule (CuPc in the case oftheir work) by a limited set of “interaction sites”representing a cluster of atoms in the molecule, andspatially arranged to approximate the shape of theactual molecule. The Lennard-Jones potential isthen used, along with effective vdW coefficientsderived for each interaction site, to calculate thelowest energy crystal structures. For CuPc, the 54atoms were replaced by 13 spherically symmetricsites, thereby eliminating 95% of the pairwise cal-culations required to determine the interaction be-tween two like molecules. The number and distri-bution of sites chosen is determined by a compromisebetween the requirement to simply represent molec-ular geometry and symmetry, and the need to includeas much of the interaction physics as possible withoutintroducing an excessive number of sites. Using thismethod, a structure resembling that of the planarstacking R form of CuPc was calculated,118 althoughthe more common herringbone â-CuPc was notfound.117 Hence, while this interaction site modelconsiderably simplifies calculations for complex mo-lecular structures, as in all such methods, one mustbe aware that the approximations can significantlyimpact the accuracy of the technique.The ellipsoidal approximation of Zhang and

Forrest72(eqs 3.3 to 3.5) has been applied to deter-mined the equilibrium structure of a NTCDA mono-layer (1000Å radius) deposited on a PTCDA “sub-strate” of equal size. The preferred orientation of theNTCDA lattice on PTCDA shown in Figure 3-12 isθmin ) 0.75( 0.05 rad, defined as the angle between

φij(θ)EP ) -R(θ)/(rij - δ(θ))6 +â(θ)exp[-γ(θ)(rij - δ(θ))]rij - δ(θ) g 0 (3.3)

φij(θ)EP ) -φo exp[-|x/η(θ)|æ(θ) - |y/κ(θ)|λ(θ)] z ) d

(3.4)

(R(θ)cosθ)2/Ra2 + (R(θ)sinθ)2/Rb

2 ) 1 (3.5)

Figure 3-12. Binding energy of a monolayer of NTCDAon a PTCDA substrate as a function of relative latticeangle, θ, between the a axes of the two surface unit cells.The vertical axis is the energy normalized to the bindingenergy per NTCDA molecule (from ref 99).

1810 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 19: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the a axes of the two unit cells. The “noise” in theplot results from rounding errors and the limited sizeof the lattices considered. The minimum energy forPTCDA/NTCDA structures is broad (as compared toPTCDA-on-graphite) due to the low spatial frequen-cies of the lattices and the large spatial extent of themolecules in the two layers. Nevertheless, the exist-ence of such an energetically favored configuration(with -3.5 meV/molecule normalized to the PTCDA/NTCDA dimer binding energy of -0.90 eV) suggeststhat this materials combination can grow as anorientationally ordered, crystalline multilayer stackunder the appropriate thermodynamic conditions, ashas been previously observed.23,52

From the foregoing results, we infer that thedependence of the orientation on molecular structureis implicit in the requirement that φinter′′ << φintra′′.The generality of this condition was tested by calcu-lating φinter′′ and φintra′′ as functions of molecular“shape” for several polyacenes in the series of ben-zene, naphthalene, pyrene, perylene and coronene

“deposited” on a PTCDA substrate. The calculatedenergy surfaces (normalized to their minimum val-ues, φo) for single benzene and coronene moleculeson a PTCDA substrate lattice are shown in Figure3-13. It is apparent that the energy minimum of thelarger molecule is extremely broad and flat ascompared to that of benzene. Hence, the shear stress(φinter′′) for the coronene/PTCDA interface is muchsmaller than that of benzene, which should enhancethe probability for successful QE growth of incom-mensurate overlayers of the larger molecule.The normalized shear stress (∆2φinter/∆a2)/φo for a

molecule on a PTCDA lattice is plotted in Figure 3-14as a function of molecular moment of inertia, Imol(normalized to benzene, Ibenz). Here, ∆2φ/∆a2 is theenergy difference divided by the surface area enclos-ing the energy 15% larger than its value at φo.Hence, ∆2φinter/∆a2 is the normalized incrementalshear stress which is numerically evaluated from thecalculated values of φ(x,y). Furthermore, the nor-malized moment of inertia of a particular molecule,

Figure 3-13. 2D potential surface of (a) benzene and (b) coronene on a PTCDA substrate. Here, the molecular planes ofthe substrate and admolecules are parallel (from ref 99).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1811

Page 20: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

Imol/Ibenz, is calculated about an axis perpendicular to,and centered in, the molecular plane of the adsorbedmolecule, providing a measure of the molecular massdistribution relevant in facially stacked, discoticsystems.The normalized shear stress decreases monotoni-

cally with increasing molecular size, or number ofcarbon rings (solid line, Figure 3-14). Hence, weexpect that the ability to grow ordered, QE layers onPTCDA (or similar planar molecule) substrates in-creases as we progress from smaller (benzene) tolarger, “rounder” molecules (coronene).Molecular shape also plays a significant role in

determining structure. Hence, Imol was also deter-mined for a series of linear polyacenes deposited onPTCDA (dashed line). Here, the calculation is madefor benzene, naphthalene, anthracene, tetracene, andpentacene (with 1, 2, 3, 4, and 5 rings, respectively).The shear stress decreases monotonically for thethree smallest molecules, at which point it begins toincrease for the larger molecules. This increaseresults since PTCDA has a perylene “core” whichextends to only three rings. Hence, longer molecules(e.g., tetracene and pentacene) extend beyond thecentral molecule in the PTCDA surface cell, andoverlap the spaces between adjacent substrate mol-ecules. For the longest molecules, the rings canextend to the carboxyl end groups of adjacent sub-strate molecules. These effects will tend either toleave the shear stress unchanged (as in the case oftetracene-PTCDA), or result in a small increase dueto core repulsion from adjacent molecules (pentacene-PTCDA).These results suggest that QE is favored when both

the molecular shape and size of the substrate andoverlayer molecules are approximately matched.This suggests that ordered QE growth is a generalproperty of a large range of planar molecules whichare bonded primarily by vdW forces to the substrate.

We emphasize that while these calculations predictthe favored, or minimum energy configuration of thegrown layers, the degree and extent of layer orderingwhich is actually achieved depends critically on thenonequilibrium thermodynamic conditions underwhich growth occurs. These conditions are notpredicted by this inherently static model, althoughwe find that certain low energy “thresholds” betweenstructural isomorphs predicted in the theory suggestthat growth at low temperatures is favorable forachieving uniform crystalline order, as has beenexperimentally observed in many cases. However,a nonstatic model is required to unambiguouslydetermine the growth conditions which lead to maxi-mum structural ordering.The molecular film alignment to the substrate also

must depend on the symmetry of the substrate. Forexample, PTCDA, NTCDA, and CuPc lattices aremonoclinic, and hence they have 2-fold symmetry. Toensure alignment of neighboring islands nucleatedseparately during the onset of growth, the substratemust have a comparable symmetry. As shown in thecase of graphite58,70 (see section 3.3.2.1), adjacentmolecular islands nucleate with their principal axesrotated at angles of m(π/3), where m ) 0, (1, (2,(3. This higher order substrate symmetry resultsin the formation of grain boundaries between islands.Other substrates with natural 2-fold symmetry (suchas an organic thin film or other monoclinic lattice),or with a preferred 2-fold directionality (e.g., glasswith a strain axis parallel to the surface,65 exposedstep edges on semiconductor substrates oriented 1-2°from a principal lattice plane, narrow polymer stripesas used as waveguide buffer layers,22,115 etc.) there-fore can serve to avoid the occurrence of such grainboundaries. Indeed, step-edge growth has beenshown119 to result in appropriate 2-fold symmetryleading to extended order in both strained vdWEsystems such as VOPc on H-terminated Si(111), andfor PTCDA on polymer ribs.22

These results suggest that large, vdW-bondedplanar molecules readily order on a variety of sub-strates without necessitating that the overlayer andsubstrate be lattice-matched or commensurate. Wenote, however, that simply depositing large planarmolecules on the substrates will not necessarilyreduce the interfacial shear stress. An additionalcondition favoring QE growth is that the shape of themolecule (or the atomic mass distribution) must besimilar to the substrate molecular shape. That is,QE growth is favored for approximately round,planar molecules deposited on similarly shaped mol-ecules in the substrate. It is somewhat less favored(although not ruled out) in the case of long, linearmolecules deposited on this same substrate structure.3.3.1.2. Coincidence Models. Although energy

minimization calculations used to determine equi-librium layer structures are physically justified, thereremain several problems with their practical applica-tion and overall predictive ability. Among the mostobvious of shortcomings of these methods are givenin the following:(1) The calculations are cumbersome. Given n

atoms in the epitaxial layer and m atoms in thesubstrate, the total number of atomic pair interac-

Figure 3-14. Incremental shear stress (1/φo∆2φ/∆a2)forseveral molecules deposited on a PTCDA substrate latticecalculated as a function of molecular moment of inertiarelative to benzene. The shear stress is normalized in eachcase to the minimum dimer potential energy, φo. Themoment of inertia is calculated about the axis perpendicu-lar to and centered in the molecular plane. The solid lineis for the molecules: 1 ) benzene, 2 ) naphthalene, 4′ )pyrene, 5′ ) perylene, and 6′ ) coronene (with 1, 2, 4, 5,and 6 carbon rings, respectively), and the dashed line isfor the molecules: 3 ) anthracene, 4 ) tetracene, and 5 )pentacene (with 1, 2, 3, 4, and 5 carbon rings, respectively)(from ref 99).

1812 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 21: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

tions which must be considered is n×m. For thecalculations to be realistic, both n and m must besufficiently large using a suitable choice of latticeboundary conditions. As noted in the previous sec-tion, for PTCDA on HOPG, ∼100 unit cells wereconsidered before adequate accuracy was achieved.There are means to make these calculations moreefficient by, for example, recognizing that vdW in-teractions fall off as 1/r6. Hence, only nearest- andnext-nearest-neighbor interactions need to be con-sidered. Alternatively, approximations such as theellipsoidal potential or interaction site methods canbe employed. Nevertheless, the method is calcula-tionally intensive with results which often do notprovide an intuitive understanding of the systemsunder study.(2) The number of degrees of freedom is very large.

Compounding the problem of large computationalscale is the high multiplicity of degrees of freedominherent in the complex molecular structures andpositions within the unit cell basis. That is, byrotating, twisting, or stretching a bond within amolecule, or by changing the relative positions of twomolecules within a unit cell, the total energy of theoverlayer with respect to the substrate can be sig-nificantly altered. An accurate, predictive calculationmust therefore include minimization of total energywith respect to all of these degrees of freedom. Asnoted in the previous section, the calculations wereconsiderably simplified (and indeed made tractable)by significantly limiting the degrees of freedomavailable to the adsorbed layer. For example, mol-ecules were restricted to lie in a single plane, andeven their internal bond lengths were, in effect, fixedby assuming an “ellipsoidal” potential approximation.(3) The calculations must include sufficient physics

to be accurate. For PTCDA on HOPG or NTCDA,only vdW bonds were included in the treatment.While this is adequate for many molecular/substratesystems, in some cases (e.g., VOPc on Si(111)),Coulombic attraction, hydrogen bonding, higher ordermultipoles, etc. must also be included.109 If Coulombforces are small, they can complicate the problemsignificantly due to their long-range nature.Hence, other methods have been suggested to

attain a more “intuitive” picture for determining thepreferred azimuthal orientation, θ, of the adsorbedlayer and the substrate lattices in incommensurate,QE systems. The most common methods can broadlybe termed “coincidence models”.43,96,104,120 These mod-els are based on two hypotheses: (i) Overlayer energyis minimized by maximizing commensurability withthe substrate. That is, the minimum energy isachieved when the largest number of lattice sites arecommensurate with substrate lattice sites. (ii) “Com-mensurability” on some arbitrarily defined latticescale can often be achieved by azimuthally rotatingone lattice with another. This commensurability,when characterized by the alignment of two highsymmetry axes of the lattices is called “point-on-linecoincidence”. An example of such coincidence isshown schematically in Figure 3-15, where com-mensurate, incommensurate and point-on-line coin-cident lattice alignments are schematically illus-trated. Note that these “models” still rely on the

primary condition for QE: i.e., φintra′′ >> φinter′′, wherecoincidence is achieved only by assuming a rigidoverlayer largely undistorted by its deposition on thesubstrate.Point-on-line (POL) coincidence has been pro-

posed43,96 to govern the orientation of such materialsas PTCDA on HOPG, where two coincident latticesare found relative to the (0001) graphite surface. InFigure 3-16, we show these two orientations corre-sponding to angles of 3.1° and 9.9° of the PTCDA(010) with the graphite symmetry axis. Both of these

Figure 3-15. Schematic diagrams of three types of epi-taxial orientations: (a) commensurate, (b) incommensu-rate, and (c) point-on-line coincident growth (from ref 123).

Figure 3-16. Schematic diagrams of two different point-on-line coincidence orientations of PTCDA on graphite(0001) proposed by Hoshino et al. (ref 43). For the type Iorientation (upper panel), PTCDA molecules at (3a,b) onthe equilibrium (bulk) surface unit mesh nearly coincidewith points on the graphite lattice. For type II orientations(lower panel), PTCDA molecules at (2a,b) are in closecoincidence with points on the graphite lattice.

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1813

Page 22: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

orientations have been experimentally inferred fromanalysis of the Moire patterns observed in the STMimages (see section 3.3.2.1, Figure 3-21), and theorientation at 9.9° has also been determined fromminimum energy calculations discussed in the previ-ous section.While this approach can provide some physical

insight into the alignments observed, coincidencemodels lack a compelling physical foundation.121Since they are not based on energy arguments, theyare incapable of predicting the preferred orientationout of many potential POL coincidence possibilities.Furthermore, the results are completely independentof the details of physical structure of the moleculeswithin a unit cell, or the degrees of freedom accessibleto the atoms within the cell. For example, if moleculeA in a lattice with a basis of M molecules is replacedby a completely different molecule, B, in a unit cellof equivalent dimension but with a basis M′, the POLtechnique will predict exactly the same overlayerorientation as in the first case. This is a naturalconsequence of the coincidence models dependenceon geometrical rather than energetic considerations.Finally, although the POL model requires only thatlattice vectors coincide in the substrate and adsorbatelattices, the molecular and atomic sites need only becommensurate at the corners of the larger “supercell”such as those outlined in Figure 3-16. More extremeexamples of such commensurate supercells120 areprovided in Figure 3-17. In such examples, there isno obvious energetic justification for requiring coin-cidence along the chosen axes, where we emphasize

that virtually all molecules and atoms in the super-cells are incommensurate with the substrate lattice.While these geometrical, POL coincidence methods

have “little direct theoretical significance”,120 Hillerand Ward have made some useful approximationswhich provide an energy-based justification for theobservation that several QE systems appear to favoralignment at, or close to substrate axes of highsymmetrical order. We first recognize that simple(i.e., monoatomic) lattice potentials can be approxi-mated by a periodic potential of the form (in onedimension):

where Vo is the strength of the interatomic potential,x is the displacement, and a is the substrate (1D)lattice constant. Defining the transformation matrix,C, which depends on the relative azimuthal relation-ship between the adsorbed layer and the substratelattice, we have

where i,j are the lattice sites of the adsorbed layers.The matrix, C, is defined by

where a1,2, R and b1,2, â are the primitive vectors andangles of the substrate and adlayer unit meshes,respectively (see Figure 3-16). The results120 ofcalculating the relative potential, V/Vo, as a functionof azimuthal angle, θ, for the several supercells ofFigure 3-17 are shown in Figure 3-18. Here, V/Vo )0 corresponds to the case of perfect commensurability(i.e., true epitaxy), whereas complete incommensu-rability corresponds to V/Vo ) 1. In general, there-fore, QE is described by systems whose energy liesbetween these extremes. In each of the cases studiedby Hiller and Ward, there is a well-defined angle atwhich the supercell achieves a minimum energy,consistent with full energy calculations discussed insection 3.3.1.1.Whether or not this method predicts the true

azimuthal orientation of the two lattices dependscritically on the choice of potential for the substrate,the shape, and composition of the overlayer mol-ecules, and the symmetry and composition of thebasis lattices. While simple vdW systems such asPTCDA on HOPG may be adapted to this treatment,the relative orientation of even such simple lattices

Figure 3-17. Schematic representations of overlayerorientations on substrates for several incommensuratesystems. The perimeter of the primitive unit mesh is shownby the lines in the corners of the larger, coincident,nonprimitive supercells exhibiting POL coincidence. Thevarious systems correspond to (a) PTCDA/Cu(100), (b)CuPc/MoS2, (c) (Pe)2ClO4/HOPG, (d) TTF-TCNQ/Au(111),(e) â-(ET)2I3 (type I)/HOPG, and (f) â-(ET)2I3 (type II)/HOPG. Here Pe ) perylene, ET ) bis(ethylenedithio-lotetrathiafulvalene), TTF ) tetrathiafulvalene, TCNQ )tetracyanoquinodimethane (from ref 120).

V ) Vo cos(2πx/a) (3.6)

V

Vo

)

1

2

∑i∑j

2 - cos[2π(ij)[C](10 )] - cos[2π(ij)[C](01 )]∑i∑j

(3.7)

C )

(b1 sin(R - θ)/a1 sin(R) b1 sin(θ)/a2 sin(R)b2 sin(R - θ - â)/a1 sin(R) b2 sin(θ + â)/a2 sin(R) )

(3.8)

1814 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 23: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

cannot be accurately predicted due to the lack ofcorrespondence between the choice of V(x) and thetrue interatomic potentials which fully incorporatesthe inherent “flexibility” of the adsorbate lattice. Thisflexibility manifests itself via the stretching, twisting,or rotation of any number of inter- and intramolecu-lar bonds which are not considered in the POLcoincidence technique. Indeed, in the POL treat-ment, each molecule in the adsorbate layer wasapproximated by a single “point” potential. Whilethis is a reasonable approximation of the case of aninert gas molecule (e.g., Ar) adsorbed on graphite, theapproximation must break down for spatially ex-tended molecules whose spatial period is large com-pared to that of the substrate, such as an OMC onan inorganic or organic substrate. Nevertheless, suchcoincidence models appear to explain observations ofsome simple vdW-bonded systems, and provide anintuitive rationale which is useful in guiding morecomputationally intensive methods such as thosediscussed in section 3.3.1.1.3.3.1.3. Mass Density Waves. It is useful to

compare the process of QE with the previouslyobserved phenomenon of “orientational epitaxy” whichdescribes the structure of layers of physisorbed inertgas atoms bonded by vdW forces onto lattice-mismatched graphite substrates. The most obviousdifference is that, in the case of orientational systems,the spatial frequency of oscillation of the substrateatomic potential is nearly equal to that of theadsorbed inert gas layer. The interatomic distanceis simply determined by the packing of the atomichard spheres. This is in contrast to QE systemswhere the adsorbed molecular spacing is substan-tially larger than the interatomic distances in thesubstrate lattice. It was proposed that mismatchbetween adsorbate inert gas atoms and the substratein orientational systems results in small, periodic,static density variations (mass density waves, orMDWs) in the overlayer.75,116 Competition betweenlongitudinal and transverse modes in these MDWs

results in an energetically preferred angle (θmin)between reciprocal lattice vectors G and τ of thesubstrate and adsorbed layers, respectively. It wasshown that, for the special case of long MDWwavelength, θmin is determined by minimizing:

where ujG2 is the average displacement of a molecule

in the overlayer,M is the molecular mass, and εl andωl are the eigenvectors and eigenvalues of the dy-namical equation defined by D(ωl)λ(ωl) ) ωlλ(ωl).Here, D(ωl), is the dynamical matrix operator,80 andλ(ωl) are the Bloch functions for the PTCDA surfacelattice. This approach requires that the phonondispersion curves (ωl(G)) for the adsorbed 2D latticesbe calculated along the various crystal axes assumingvdW bonding described by eq 3.2. When applied toPTCDA and NTCDA, these calculations yield twotransverse and one longitudinal acoustic phononmode.99 Optical modes were ignored since ωl forthese terms is large, resulting in a small contributionto eq 3.9.Summing the contributions from each mode gives

the MDW energy for NTCDA shown in Figure 3-19.The several peaks correspond to the summation ofphonon modes, ωl, along the various crystal direc-tions, with an energy minimum at θmin ) 0.65 rad.Note that the contribution to the total layer energydue to this lattice distortion should be small inincompressible molecular lattices (whereMωl

2 is verylarge due to the large molecular mass and highvibrational energy), and thus should not significantlyalter the position of the NTCDA/PTCDA energyminimum from that shown in Figure 3-12.Note that the lattice distortion energy accounted

for in eq 3.9 is a second-order effect in molecularsystems, whereas in relatively compressible atomicvdW lattices (where M and ωl are small) this effectshould dominate.75,116 However, the presence ofMDWs in QE systems may nevertheless provide amechanism for reduction in total strain energy, and

Figure 3-18. Normalized energy (V/Vo) on azimuthal angle(θ) for the overlayer-substrate combinations in Figure3-16, assuming an overlayer dimension of 20 × 20 unitcells: (a) PTCDA/Cu(100), (b) CuPc/MoS2, (c) (Pe)2ClO4/HOPG, (d) TTF-TCNQ/Au(111), (e) â-(ET)2I3 (type I)/HOPG, and (f) â-(ET)2I3 (type II)/HOPG (from ref 120).

Figure 3-19. Relative lattice energy for a NTCDA surfaceunit cell on NTCDA as a function of orientation, θ, as aresult of strain induced in the adlayer. This energy is inaddition to the rigid lattice energy calculated in Figure 3-12and is expected to be small by comparison (from ref 99).

φ ) -∑l

ujG2 [G‚εl(G)]

2

Mωl2(G)

(3.9)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1815

Page 24: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

indeed may already have been observed for PTCDAon HOPG (using STM) and on Au(111) (using X-raydiffraction).58,73 That is, MDWs may manifest them-selves as periodic variations in a degree of freedomwith a unit cell (e.g., a small angular rotation of themolecules with respect to each other, with the angleperiodically varying over several unit cells). Theseperiodic variations may provide a means to reducethe total crystal energy induced by lattice strain,thereby eliminating the formation of dislocations orother macroscopic lattice defects. Furthermore, theaddition of the MDW energy to that of the otherwiseunperturbed adlayer via eqs 3-9 may induce somedeviation from point-on-line reconstruction.

3.3.2. Observation of QE Growth

To test the above assumptions regarding the condi-tions which are required to achieve QE thin films,PTCDA has been grown on several different sub-strates with varying magnitudes of substrate/adlayerinteractions. The case of PTCDA on HOPG has beenthe most thoroughly studied, and from the foregoingdiscussion, it is apparent that the QE layer shouldbe azimuthally oriented with respect to the graphitelattice without inducing significant strain in thethicker QE layer. In addition, PTCDA has beengrown on Au(111) and GaAs which have somewhatstronger substrate-molecular interactions than inthe case of HOPG, resulting in a strained-layer QEPTCDA thin film. Below, we consider each of thesecases in detail to provide a comprehensive picture ofthe nature of QE thin-film growth for different bondenergies between the substrate and epitaxial layer.3.3.2.1. STM Imaging of PTCDA and NTCDA

Monolayers on HOPG. The onset of QE growth hasbeen thoroughly investigated by examining mono-layers of PTCDA43,57,58,70 and NTCDA122 grown onHOPG under UHV and atmospheric ambient condi-tions. Experiments described here were primarilyconducted in UHV, using electrochemically etchedtungsten tips and mechanically sheared Pt-Ir tips.58The PTCDA was deposited at a rate of 3-10 Å/minon substrates held from 80 K to slightly above roomtemperature,43,58,70,122 whereas the NTCDA growthswere performed for substrate temperatures in therange of from -30 to 100 °C.122 The film thicknessstudied in these experiments ranged from submono-layer coverage up to 2-3 ML. While it is difficult toobtain STM images with molecular resolution forthicker layers due to the low conductivity typical ofOMCs, nevertheless, PTCDA layers up to 60 Å thickdeposited on Se-passivated (100)GaAs have beensuccessfully imaged using UHV STM.123

Figure 3-20 is a high-resolution image of a 1 MLthick film of PTCDA. The shapes of the two mol-ecules within the unit cell are distinct; one appearsas a figure-eight while the other less well resolvedmolecule appears as two parallel lines. The dimen-sions of the PTCDA rectangular unit cell (Figure3-20) were found to be a ) 13.2 ( 0.45 Å and b )19.5 ( 0.6 Å, which agree with measurements ofPTCDA on HOPG43,57,70 done under ambient condi-

tions suggesting the as-grown layers are stable.These results are also consistent with the enlarge-ment of the bulk unit cell36 in the (102) plane (a )11.96 Å, b ) 19.91 Å). A similar cell enlargementhas been observed the first few ML of PTCDA onAu(111)73 and for PTCDA on alkali halide sub-strates.46,124 As noted in section 3.3.1.1, this enlarge-ment is anticipated for monolayer QE films ofPTCDA due to the reduced vdW bond energy in 2Das compared with 3D.The orientation of the molecular layer with respect

to the HOPG can be obtained by maneuvering thetip into holes in the overlayer58 or by removing thelayer with the tip.122 Alternatively, by increasing thecurrent, the tip penetrates the thin film, providing adirect image of the underlying graphite lattice.70 Bythese means, the angular relationship between thesubstrate and the overlayer can be directly deter-mined. For PTCDA, two domain orientations areobserved with respect to the HOPG: The first (typeI) is at an angle of 3° ( 3° between the a axis ([010]direction) of the PTCDA and the [112h0] direction ofgraphite,43,57,58 and the second1,43,58,70,123 (type II) isat an angle of 10 ( 1°. Although these measure-ments are only accurate within 1° or 2°, the type IIorientation is clearly consistent with the calculatedstructure, which predicts θ ) 10.7 ( 1°. Botharrangements are consistent with the POL coinci-dence fits of 3.1° and 9.9° (see Figure 3-16). All ofthese experimental and theoretical results obtainedfor PTCDA on HOPG are summarized in Table 4.Moire patterns shown in Figure 3-21 indicate a

periodic superstructure at an angle of 10.5° withrespect to the b axis ([2h01] direction) of the PTCDAunit cell. This result differs from Ludwig et al.42 inwhich Moire fringes were observed parallel to the baxis resulting from a proposed commensurate 1×3superstructure. Due to the weak vdW bond and the

Figure 3-20. A 200 Å × 200 Å STM image of PTCDA onHOPG showing a unit cell consisting of two moleculesindicating the crystallographic directions. Image taken inthe constant current mode under UHV conditions, with atip current of 200 nA and voltage of -800 mV (from ref58).

1816 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 25: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

numerous degrees of freedom available to the layeras a means for reducing total energy (c.f., section3.3.1.1), it is unlikely that the molecules form a rigid,commensurate reconstruction on a scale equivalentto the repeat distance in the Moire pattern (∼119 Åalong the [010] direction and∼215 Å along the ([2h01]direction). Indeed, such a commensurate superstruc-ture (extending over >32 substrate unit cells) hasalso been proposed122 for NTCDA on HOPG, andNTCDA on MoS2, which is equally unlikely, particu-larly in view of the fact that small inaccuracies incalculating the size of the overlayer cell from the STMimage can lead to large variations in this superstruc-ture periodicity. Instead, these results imply that theoverlayer is incommensurate with the substrate,consistent with the QE growth model.To further understand the contrast apparent in

STM images of organic monolayers, images have beenobtained for a wide range of tip biases. For example,a voltage-dependent contrast of the PTCDA layer hasbeen observed.58 Recall that the two molecules in the

bulk (102) plane are tilted out of the plane,36 asituation which might also obtain for monolayer filmson HOPG. The slight difference in height betweenthe two molecules would be observable when the tipis closest to the film (i.e., low bias voltages).However, the surface unit cell is expanded from

that of the bulk, suggesting that molecules in the firstML probably lie parallel to the substrate plane.Thus, the molecules aligned with their long axis closeto a symmetry axis of the substrate are expected tohave low contrast due to increased interaction withthe substrate. In this case, the molecules imagedwith higher intensity at low biases would actuallybe less strongly bound due to a reduced substrate/adlayer orbital overlap. At higher biases, thesesubtle effects would not be observed,58 giving rise toan equal intensity for both molecules in the STMimages as observed. It is likely that a combinationof both topographic and electronic effects contributeto the voltage dependent contrast. Indeed, thisinterpretation has been confirmed by electron-densitycalculations for the PTCDA/graphite interface.125

Figures 3-22a and 3-23a are STM images of filledand empty states, respectively. The images are takenat the lowest bias voltages below which the contrastof the organic monolayer is unobservable. Underthese conditions, it is found that the PTCDA mol-ecules can be imaged with submolecular resolution.58Molecular electron density calculations based onindependent optimization of molecular orbitals (orZINDO)43,58,126-128 are shown in Figures 3-22b and3-23b for PTCDA. There is a clear correspondencebetween the calculated HOMO and LUMO bands andthe filled and empty electronic state images, respec-tively. The highly electronegative oxygen atoms inthe carbonyl groups are calculated to have an excesscharge of ∼0.4 e compared with that on the nearestcarbon neighbor. Nevertheless, the energy of thesestates lie lower than the HOMO, with no significantcontribution to the STM images. Instead, the πbonds in the perylene core contribute most to deter-mining the HOMO energy, and hence are mostclearly apparent in the STM images.The interpretation of these molecular images has

been tested by Strohmaier et al.122 who also examined

Table 4. Summary of Growth Parameters Measured for PTCDA on HOPGa

analytical techniqueunit cell

dimensions (Å)orientation to[112h0] (deg)

molecular axesorientation

Bulk PTCDA (102) plane (from XRD) (ref 36) a ) 11.96 48.9° to [2h01]b ) 19.91 41.1° to [010]

STM (refs 1 and 70) a ) 15.2 ( 1.6 θ ) 13.0 ( 4 55 ( 1.7° to [2h01]b ) 21.6 ( 2.2 35 ( 1.7° to [010]

total energy calculation (refs 72 and 99) a ) 15.6 ( 0.5 49.3 ( 1° to [2h01]b ) 19.8 ( 0.5 θ ) 10.7 ( 1 40.7 ( 1° to [010]

LEED (ref 42) a ) 12.5 ( 0.75 θ ) -3 ( 0.15b ) 19.2 ( 0.7

STM (ref 42) a ) 12.7 θ ) (3.2 ∼49.5 to [2h01]b ) 19.2 ∼40.5 to [010]

STM (ref 43)polytype I a ) 12.69 θ ) (3.1

b ) 19.22polytype II a ) 12.37 θ ) (9.9

b ) 19.43STM (ref 123) a ) 13.2 ( 0.45 θ1 ) (3 48.6 ( 1.9° to [2h01]

b ) 19.5 ( 0.6 θ2 ) (10.5 ( 3 41.4 ( 1.9° to [010]a Table from ref 123.

Figure 3-21. A 200 Å × 200 Å STM image of a monolayerof PTCDA on graphite showing Moire fringes orientedparallel to the long arrow. The fringes are at a 10.5° angleto the b axis (at [010]) of the PTCDA unit cell (from ref58).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1817

Page 26: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the smaller molecule, NTCDA. If the contrast mech-anism is the same for both PTCDA and NTCDA, thenumber of contrast maxima should be equal to thenumber of C-C bonds in the molecular core, corre-sponding to the ten observed for PTCDA and six forNTCDA. This has been confirmed as illustrated inthe STM image of Figure 3-24a where the “figureeight” characteristic of the larger PTCDA moleculehas been replaced by a simple “ringlike” image forNTCDA. In Figure 3-24b, molecular orbital calcula-tions of the NTCDA LUMO are shown which cor-respond well to the observed image of the C-C bondarrangement.Since the contrast in the STM images have been

confirmed to result from the overlap of molecularorbitals in the thin film and substrate, the imagestriations, or Moire patterns, can thus be analyzed

to accurately determine the relative orientation of thetwo lattices.96 This can be understood by consideringthe geometrical relationship between the two con-tacting lattices in Figure 3-25. Here, we write theunit vectors of the molecular layer, a and b in termsof the substrate vectors, ag and bg using

where i, j, k, and l are integers, and ∆a and ∆b arevectors which arise due to the mismatch between theincommensurate lattices. It can be shown thatdepending on the values of ∆a and ∆b, either a twodimensional or one dimensional Moire pattern willresult. These two types of patterns are illustrated

Figure 3-22. (a) Low bias, 100 Å × 100 Å filled state STM image of PTCDA on graphite. Measurement conditions are atip current of 0.1 nA and a voltage of -250 mV. (b) Molecular orbit calculation indicating the HOMO of a PTCDA molecule(from ref 58).

Figure 3-23. (a) Low bias, 100 Å × 100 Å empty state STM image of PTCDA on graphite. Measurement conditions area tip current of 0.1 nA and a voltage of +250 mV. (b) Molecular orbit calculation indicating the LUMO of a PTCDAmolecule(from ref 58).

a ) kag + lbg + ∆a (3.10a)

b ) iag + jbg + ∆b (3.10b)

1818 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 27: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

in Figure 3-26, parts a and b, respectively, and haveboth been observed experimentally (see Figure 3-21for a 1D pattern58 resulting from PTCDA on HOPG,

and Figure 3-27 for a 1D image96 due to 1,4-dithioketo-3,6-diphenylpyrrolo[3,4-c]pyrrole, or DTPP, on HOPG).With a coordinate transformation, vectors ∆a and ∆bcan be rewritten in terms of the substrate latticecoordinates using

where p, q, r, and s are numbers. Clearly, for thespecial case of ∆a|∆b, these values are related by ps- qr ) 0. Referring once again to Figure 3-26, theMoire fringe unit vectors can be expressed in termsof the transformation coordinates, a′ and b′, byintroducing an additional set of numbers, λa, and λb,which provide the periods of the contrast modulationalong these two directions. Hence, the Moire pat-terns become periodic (with well-defined “basis” vec-tors, aM, bM) for

where integers m and n satisfy the relationship pn- qm ) 0. For a 1D pattern, the equation of theMoire line intersecting the origin is λay + λbx ) 0.Furthermore, the existence of these patterns implythat ∆a and ∆b are parallel to the substrate direc-tions defined by the lines (n,-m), and the end pointsof the vectors a′ and b′ lie on that same line.From this treatment, an accurate relationship

between the ML film and the substrate lattices canbe determined from the existence of the Moire pat-terns in the STM image. This has been done byHoshino and co-workers43,96,104 for the cases of PTC-DA, DPP, and other molecules deposited on HOPG.For PTCDA on HOPG, the results obtained from theMoire analysis are consistent with earlier, directmeasurements where the substrate and film wereindependently imaged (see Table 4). We note, how-ever, that the Moire analysis is considerably moreaccurate, providing angular relationships to perhapswithin (0.5°, whereas direct imaging accuraciestypically do not exceed (2°.The Moire fringe patterns therefore provide a

striking example of several predicted features of QEgrowth. For example, the patterns may indicate thepresence of MDWs acting to reduce stress due tolattice mismatch (see section 3.3.1.3), although thisexplanation is difficult to verify given the limitedspatial resolution of the STM image. Alternatively,the Moire pattern may be due to electronic orbitalinteractions between the film and the substrate, ashave been observed by STM for the vdWE growth oftransition metal dichalcogenides69,129 such as MoSe2on MoS2. A small angular misalignment of the filmproduces a large change in the angle of the Moirepattern. Nevertheless, whether the observed con-trast modulation is due to MDWs or due to orbitaloverlap (i.e., electronic) effects, both explanations areconsistent with an incommensurate overlayer havinga structure which, to first order, does not depend onthe lattice constant of the substrate, but which tends

Figure 3-24. (a) STM image of NTCDA molecules ongraphite showing six tunneling current maxima in ahexagonal arrangement corresponding well to the LUMOcalculated for this molecule shown in b. Imaging conditionswere a sample voltage of -900 meV and a tip current of45 pA (from ref 122).

Figure 3-25. Schematic of the surface unit mesh of anorganic epitaxial layer (solid lines) with dimensions a, b,lying on a substrate (dashed lines) with unit mesh param-eters, ag, bg. Also shown are the “discrepancy vectors”, ∆aand ∆b and reduced lattice vectors, a′ and b′ (from ref 96).

a′ ) ag + ∆a ) ag + pbg + qbg (3.11a)

b′ ) bg + ∆b ) bg + rag + sbg (3.11b)

∆a ) (mag + nbg)/λa (3.12a)

∆b ) (mag + nbg)/λb (3.12b)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1819

Page 28: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

to slightly misalign from a symmetry axis in orderto reduce strain.72Scanning tunneling microscope images taken over

larger fields of view show boundaries between adja-cent OMC film domains. For example, an STMimage over an area of (750Å)2 near such a domainboundary of PTCDA on HOPG is shown in Figure3-28a, with a similar image for NTCDA on HOPGshown in Figure 3-28b. In both cases, the anglebetween the epitaxial molecular axes measured indifferent grains is always observed to be an integermultiple of π/3 rad.1,43,58,122 Although HOPG pos-sesses only 3-fold rotational symmetry, nucleatingislands of PTCDA on the surface of graphite areenergetically identical at rotations of π/3 rad.99 Theabsence of any other angles in the observed STMimages provide additional strong evidence for apreferred angle of orientation between the organiclayer and the substrate, which is striking evidencefor QE growth. That is, even though the two latticesare incommensurate, they align in a consistentfashion with respect to each other. This observation

also points to a fundamental and obvious limitationof quasiepitaxy: If the symmetries of the two latticesare not identical, the deposited layer must containhigh-angle grain boundaries between islands whichhave separately nucleated at different lattice sites.Given that the monoclinic cells characteristic of manycrystalline organic semiconductors have only 2-foldsymmetry, achieving perfect or near-perfect filmorder is restricted to growth on substrates withsimilarly low symmetries.65,102,119Finally, STM has also been used to determine the

relative orientation between layers in PTCDA filmsof from 2 to 3 ML in thickness in order to study theevolution of film growth immediately following nucle-ation.58 Figure 3-29 shows detailed images of twosides of a PTCDA step from 1 to 2 ML coverage. Thedomain orientation is clearly the same in each of thetwo layers, with the step-edge aligned along thePTCDA a axis. The observation of preferentialalignment of step-edges along the a axis implies ananisotropy of the intralayer binding forces which aregreater perpendicular to the b axis than to the a axis,leading to low-energy steps along the [010] direction.That is, the lack of small islands suggests that thereis high surface molecular mobility even at roomtemperature and that molecules have a significantlylower energy when attached to a large domain.Furthermore, these STM images are consistent

with QE ordering. Specifically, growth begins as anordered overlayer which is not lattice matched to thesubstrate and continues in a quasi-epitaxial, slightlystrained, layer-by-layer manner.70,73,74,99 In addition,the small angle observed between the PTCDA andthe [112h0]HOPG axis is consistent with calculationswhich suggest that the overlayer orientation isdetermined by energy minimization with respect tothe substrate. Due to the low energy of the vdWbond, however, the strain induced by the latticemismatch is not sufficient to strongly distort theoverlayer on a scale observable by STM.3.3.2.2. In Situ Observation of the Evolution of QE

Growth. RHEED has proven to be a useful tool forproviding a direct observation of the evolution ofPTCDA film nucleation and subsequent growth frommonolayer (ML) to multilayer stacks.52,100 Use ofRHEED as a means for in situ growth has led to amore complete understanding of how highly ordered

Figure 3-26. Calculated Moire patterns for two representative cases of an organic layer on a substrate following a point-on-line coincidence behavior. The larger points indicate where the center of a molecule lies closest to a substrate latticepoint. The diagrams correspond to (a) a′ ) 1.1ag - 0.02bg, b′ ) 0.01ag + 1.15bg. The small parallelogram indicates theunit cell of the adsorbed layer, and aM and bM are the unit cells of the Moire pattern. (b) a′ ) 1.1ag, b′ ) -0.08ag + bg.Here r| indicates the direction of the one dimensional Moire line, and λa, λb are the periods of the 1D contrast modulationalong a′ and b′ in the adsorbed layer, respectively (from ref 96).

Figure 3-27. A 250 Å × 250 Å filtered STM image ofDTPP film deposited on graphite. The parallelogram in theimage indicates the surface unit mesh of DTPP. 1D periodiccontrast modulation of the Moire pattern is indicated bythe arrows (from ref 96).

1820 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 29: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

QE and vdWE films can be achieved.2,3,81,82,91,97,105,130-135

Typically, RHEED patterns are obtained using an ∼8keV electron beam directed onto the film surface atan angle of from 1° to 2°. Care must be taken tominimize exposure of the thin film to the e beamsince extended bombardment by high energy elec-

trons can cause damage to fragile organic moleculessuch as NTCDA.100

A typical series of RHEED patterns1,52 obtainedduring a growth sequence of PTCDA on HOPG isshown in Figure 3-30. The RHEED pattern for thebare graphite substrate is shown at the top left.From the measured spacing of the streaks and thebeam parameters, a surface unit cell (or surface net)spacing of 2.17 ( 0.04 Å was measured, correspond-ing to the (220) reflection of graphite. After ∼1 Å ofPTCDA is indicated on a crystal microbalance thick-ness monitor, the graphite streaks fade, and com-pletely vanish after a full ML (∼3 Å) is grown. At afilm thickness of 10 Å (∼3 ML), the graphite featuresare replaced by a set of long, continuous streaksshown in Figure 3-30, indicating that the first fewlayers of PTCDA are both crystalline and smooth.The dimensions of the surface unit mesh inferredfrom these patterns are compiled in Table 3 andindicate a somewhat expanded surface reconstructionfrom the bulk, consistent with both calculations andSTM data also provided in the table.Continuous streaks indicate film flatness (to within

a monolayer) over a coherence length of the electronbeam.136,137 While the contact between the beam andthe film surface is typically a few millimeters inlength, the coherence length is only on the order of afew hundred Ångstroms. Hence, continuous streaksindicate a molecularly smooth surface over at leastthis distance. STM studies discussed in the previoussection have indicated a layer-by-layer growth modeon HOPG covering areas g(1000 Å).2 As will beshown in the following section, this distance in factexceeds several micrometers for PTCDA growth onAu(111) which has a substantially higher interactionenergy (and correspondingly larger φinter′′) than forHOPG. Hence, we anticipate that continuous RHEEDstreaks in the case of PTCDA and NTCDA ongraphite are indicative of molecularly flat layers overextremely large distances.

Figure 3-28. (a) STM image of a domain boundary ofPTCDA grown on a graphite lattice. The arrows in eachdomain indicate the b axis of the PTCDA lattice and thedirection of the 1D contrast modulation. Experimentalconditions are a tip current of 20 pA and a sample voltageof 400 mV (from ref 43). (b) A 500 Å × 500 Å STM imageof a domain boundary of NTCDA grown on graphite. Theangle between domains is 120° due to the hexagonalsymmetry of the substrate. Experimental conditions are atip current of 18 pA and a sample voltage of-500 mV (fromref 122).

Figure 3-29. STM image of a step up from a monolayerto a second layer of PTCDA deposited on graphite. Thestriations in the image are due to heightened contrast alongthe a axis of the PTCDA unit cell to which the step isaligned (from ref 58).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1821

Page 30: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

As growth proceeds to 20 and 50 Å, the positionsof the intense streaks remain unchanged, althoughvariations in intensity begin to appear at 50 Å (∼15ML) along the length of the streaks, even though thegrowth rate has remained at 0.5 Å/s. These RHEEDpatterns show that the crystalline organic layerretains a high degree of surface planarity and order.As the film thickness is increased to 220 Å, and thenfinally to 1000 Å (Figure 3-30, bottom), the RHEED

pattern periodicity remains unchanged from its valuefor thinner layers, although the continuous streakshave broken into several segments separated by veryfaint streak remnants. As noted above, this streaksegmentation indicates that the film surface hasbecome uneven on a molecular scale, although thecrystalline structure remains unchanged.The azimuthal orientation of PTCDA on HOPG has

been further examined using low energy (∼15 eV)electron diffraction which is performed at near nor-mal electron beam incidence.101 Figure 3-31, partsa and b, are LEED patterns obtained from bareHOPG and from a 50 Å PTCDA layer grown on aroom temperature HOPG surface, respectively. Bothexhibit a ring-like structure characteristic of multiplyrotated, single-crystal domains distributed on a flatsurface. In Figure 3-31a, the radius of the inner-ringcorresponds to reflections from lattice planes spacedat 2.46 Å, characteristic of HOPG. The pattern,therefore, results from a superposition of multiplehexagonal grains slightly rotated with respect to eachother.

Figure 3-30. A series of RHEED patterns obtained fordifferent thicknesses of PTCDA grown on HOPG (whosepattern is shown in the upper left). The 20 Å film RHEEDpattern shown at the bottom has been contrast enhancedto exhibit the detail in the brightest and faintest part ofthe pattern (from ref 1).

Figure 3-31. Low-energy electron diffraction pattern from(a) a bare HOPG surface and (b) a 50 Å thick PTCDA filmgrown on HOPG. The electron beam energies were 100 eVfor image a and 13 eV for image b (from ref 101).

1822 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 31: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

The PTCDA diffraction rings in Figure 3-31b arecontinuous, indicating the existence of multiple PTC-DA domains distributed inside the HOPG domains.As in the case of the RHEED experiments, thedimensions of the PTCDA single-crystal domains(approximately several micrometers) are larger thanthe LEED coherence length (100-200 Å), as evi-denced by the thinness of the diffraction rings.As discussed in section 3.3.1, QE is a kinetically

controlled growth mode, with the resulting structuredepending on a balance between thermal energy,bond energy, and rate of deposition. Hence, weanticipate that structure depends critically on thesubstrate temperature.45,47,48,138-140 While substratetemperature effects have been examined in detail forPTCDA on Au(111) (see section 3.3.2.3), they havealso been studied for PTCDA on HOPG using in situRHEED.52 In Figure 3-32, the RHEED patterns fortwo, 30 Å thick (∼9 ML) PTCDA films grown onHOPG substrates temperatures of Tsub ) 100 K (topimage), and at Tsub ) 325 K (bottom image) arecompared. The discontinuous RHEED streaks for

growth at elevated temperatures suggests a some-what less planar PTCDA surface than for the caseof the low temperature growth. This difference insurface morphology is also apparent from scanningelectron microscope images taken for samples grownunder similar conditions to those in Figure 3-32. Forexample, 0.4 µm thick PTCDA films deposited atroom temperature at a rate of 3-5 Å/s consist of amesh of crystalline filaments approximately 0.2 µmlong, as shown in Figure 3-33a. These crystallinefilaments are oriented randomly on the surface. Forthe same films deposited at 90 K (Figure 3-33b), thesurface is smooth, and a uniform, more ordered, filmis achieved.Studies of QE growth of single organic layers

has also been extended to organic multilayerstructures.1,100,138,139,141-145 As will be shown in sec-tion 4, such organic multilayers have recently becomethe focus of considerable interest due to their numer-ous device applications,24-27,146-149 as well as theirusefulness in revealing many aspects of fundamentalexcitations in OMCs which have been inaccessible via

Figure 3-32. RHEED patterns obtained for 30 Å thickPTCDA films grown on a HOPG substrate held at Tsub )100 K (top) and Tsub ) 325 K (bottom) during growth (fromref 1).

Figure 3-33. Scanning electron microscope images of thesurface of a PTCDA film grown on glass at a substratetemperature of (a) Tsub ) 293 K and (b) Tsub ) 90 K (fromref 1).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1823

Page 32: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

studies of the bulk material properties.23,24,29,34,150-152

Hence, understanding of structure is integral torevealing the physical properties of these interestingstructures. Using RHEED diffraction patterns fromalternating thin films of PTCDA on NTCDA, it wasfound that each layer is crystalline with its own well-defined surface unit cell, regardless of significantlattice mismatch with the underlying layer on whichit was deposited.100

A typical series of RHEED patterns1,100 obtainedduring the growth of a PTCDA/NTCDA alternatinglayer structure is shown in Figure 3-34. Figure 3-34ashows the pattern resulting from the subsequentdeposition of 20 Å (∼3 ML) of NTCDA on 20 Å ofPTCDA (c.f., Figure 3-30) on HOPG, followed by asecond 20 Å thick layer of PTCDA; i.e., a total of 60Å of organic material. The electron beam wasobserved to damage the fragile NTCDA layer veryrapidly; the streaks fading in only a few seconds. Toensure a good surface for the growth of the secondlayer of PTCDA, therefore, the first layer of NTCDAwas not exposed to the beam. Streaks for the topPTCDA layer are apparent as in the case for a single

layer PTCDA film, and are measured within experi-mental error to be at identical lattice spacings tothose measured from the first PTCDA layer depositedon HOPG. Clearly, the second layer of PTCDAremains crystalline and two-dimensional. Further-more, the surface unit cell of PTCDA grown onNTCDA is observed to be identical to PTCDA grownon HOPG, despite the different lattices of the under-lying layers in those two cases. The RHEED patternfrom the second layer of PTCDA also indicates thatthe underlying NTCDA layer is planar, thus resultingin a sharp PTCDA/NTCDA interface.

Figure 3-34b shows the RHEED pattern resultingfrom the subsequent deposition of a second 20 Å thicklayer of NTCDA; a total of 80 Å of organic filmmaterial. As in the case of the other layers, thefourth layer of the organic heterostructure exhibitscrystalline order, with a different RHEED pattern,and hence surface unit cell, to that of PTCDA.Following the example of PTCDA, some surfacereconstruction from the bulk lattice spacings isexpected. Assignment of the streaks suggested areconstructed NTCDA surface unit cell of dimensions13.0 ( 0.6 Å × 5.1 ( 0.2 Å.

It is also apparent that the herringbone crystalstructure of NTCDA63,122,153 will result in some dif-fraction out of the substrate plane even if the layeris completely flat, i.e., the electron beam will providelimited 3D structural information along the NTCDAa axis which will manifest itself as diffuse spots inthe diffraction image. This represents a RHEEDpattern in which spots from the “3D” diffraction andstreaks from the “2D” diffraction are expected toappear in the same image. The planarity of theNTCDA layer is inferred by the ability to grow ahighly planar layer of PTCDA on its surface.

As an additional example of QE deposition, Figure3-35a shows the image from a 20 Å thick layer ofCuPc grown on top of a 20 Å thick PTCDA film onHOPG.1 The many orders of diffraction visible in thepattern indicate a crystalline layer of CuPc of unitmesh dimensions of 13.3 ( 1 Å × 17.5 ( 1 Å. Thisis reasonably consistent with the “rowlike” phasepreviously observed in STM images154,155 of a mono-layer of CuPc with a 2D surface unit cell with b )17.5 ( 1 Å, c ) 15.5 ( 1 Å, and angle between the band c axes of R ) 98 ( 1°.

Figure 3-35b shows the RHEED pattern resultingfrom the deposition of a further 20 Å of PTCDA ontop of the CuPc. The diffuse arcs indicate a layeredstructure, but no crystalline order within the layers.The CuPc surface is clearly unsuitable for growth ofPTCDA due to 3D growth in the CuPc layer, asindicated by the lack of extended streaks in the CuPcRHEED pattern. From these observations, we con-clude that although lattice matching is not strictlyrequired, a molecularly flat surface is a necessarycondition to achieve QE growth.

This apparent “asymmetry” in growth order hasalso been observed for VOPc/PTCDA multilayerstructures. It was found that,152 starting with growth

Figure 3-34. RHEED pattern of multilayers of PTCDAand NTCDA grown on a HOPG substrate. (a) PTCDA onNTCDA on PTCDA and (b) NTCDA on PTCDA on NTCDAon PTCDA. All layers are 20 Å thick(from ref 1).

1824 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 33: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

on a quartz substrate maintained at room tempera-ture, layers of PTCDA grown on VOPc exhibitedconsiderable stacking order, whereas the oppositecombination (VOPc on PTCDA) tended to result inthe amorphous growth of VOPc on an otherwiseordered PTCDA layer. This is apparent from theX-ray diffraction patterns shown in Figure 3-36a.Curve a is for a 15 nm VOPc film, whereas b is for aPTCDA film of similar thickness. The broad peakat 2θ ) 21° is the background from the substrate.For samples consisting of 15 nm of PTCDA grownon 15 nm of VOPc (curve c), peaks are due to stackingof phase II VOPc molecules along the (010) direction(at θ ) 7.54°). However, when the deposition orderis reversed, the (010) peak disappears indicatingsignificant film disorder, similar to the case ofPTCDA grown on CuPc.1

The structural differences between the films alsoinfluences their exciton spectrum, as shown in Figure3-36b. Small blue shifts of the PTCDA peak (markedby an arrow in curves b and c) are observed whenthis material is deposited on the surface of VOPc.Otherwise, the spectra of both PTCDA and VOPc

remain unchanged from their bulk absorption. Aswill be shown in section 4, this small blue shift isprobably a result of quantum size effects observedin ultrathin PTCDA layers. In contrast to theserelatively minor effects, there are considerable dif-ferences in the absorption spectra of VOPc dependingon whether PTCDA is deposited as the first or thesecond layer. These shifts, particularly in curves dand e, where the VOPc lies on the surface of thePTCDA layer, are strongly indicative of considerablefilm disorder, consistent with the X-ray data. Fromthese crystallographic and absorption data, we con-clude that the structure of the VOPc depends criti-cally on the order of growth. This “interface asym-metry” has been widely observed in the growth ofIII-V heterojunctions,156,157 and is attributed to thedifferent rates of atomic diffusion (i.e., P-As ex-change is considerably more active at InP/InGaAsinterfaces than at InGaAs/InP heterojunctions158).This observation for Pc/PTCDA interfaces also indi-cates that intermolecular diffusion can play a rolein determining film structure. It is a direct result ofthe bonding energy at the VOPc/PTCDA interface

Figure 3-35. RHEED pattern of multilayers of PTCDAand CuPc grown on a HOPG substrate: (a) CuPc onPTCDA and (b) PTCDA on CuPc on PTCDA. All layers are20 Å thick. Note that b indicates amorphous growth of thetop layer (from ref 1).

Figure 3-36. (a) X-ray diffraction data for films depositedat room temperature: (curve a) 15 nm VOPc, (curve b) 15nm PTCDA, (curve c) 15 nm PTCDA/15 nm VOPc, (curved) 15 nm VOPc/15 nm PTCDA bilayer, (curve e) 105 nmVOPc/15 nm PTCDA bilayer. (b) Optical absorption spectrafor the same films shown in part a (from ref 152).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1825

Page 34: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

and the substrate temperature during growth whichpromotes molecular diffusion along the surface. Suchdiffusion processes play even a greater role in dis-rupting the flatness of organic/inorganic semiconduc-tor and organic/metal superlattices.144,145,159,160 Thatis, the large heat of condensation associated withinorganic materials may induce significant interfacediffusion prohibiting the realization of molecularabruptness in these cases.3.3.2.3. Strain Evolution in QE Systems: PTCDA

on Au(111). An interesting feature of QE is that thestrained layer does not necessarily relax into itsequilibrium (bulk) crystal habit even after severalhundreds, or even thousands of Ångstroms have beengrown. To develop a complete picture of the natureof these thin films, nucleation, growth, and strain ofPTCDA on Au(111) have been studied73,74,140 usingglancing incidence X-ray diffraction (GIXD). Here,Au(111) serves as an ideal substrate on which tostudy strained QE films since the PTCDA/Au(111)bond energy is higher than for the relatively un-strained case of PTCDA on HOPG. However, thebinding energy is not so high that lattice-matchedepitaxy occurs. As we have shown in section 3.2,epitaxy results in the buildup of strain energy,leading to amorphous or polycrystalline growth afteronly a few monolayers are deposited. Even thoughthe substrate-molecule interactions in PTCDA/Au(111)are sufficiently strong to introduce strain into thethin film, they are not strong enough to lift thecharacteristic 23×x3 reconstruction of the substratesurface. This indicates that PTCDA molecules arephysisorbed onto Au(111) via vdW interactions andthat there is relatively little charge exchange acrossthe interface.Studies of PTCDA on Au(111) have provided the

most comprehensive and precise picture of organicthin film growth by either epitaxial or quasi-epitaxialmodes. For these studies,73,111 PTCDA films weregrown on Au(111) substrates by OMBD at a rate of0.1-0.2 Å/s. The growth was performed in the samechamber which housed the X-ray diffracto-meter (with an X-ray wavelength of λ ) 1.285 Å)attached to the National Synchrotron Light Sourceat Brookhaven National Laboratories.The 2D structure of PTCDA/Au(111) grown at room

temperature can be determined from the Bragg peakpositions as a function of momentum transfer parallelto the surface, Q|, and the azimuthal orientation, φ,of the diffraction peaks with respect to the Au surfaceBragg rods. The spacing and stacking of the 2DPTCDA lamella were determined by the magnitudeof the momentum transfer perpendicular to the

surface plane, Qz. Table 5 summarizes the GIXDdata taken from the diffraction spectra in Figure3-37. The lattice spacings are consistent with thoseobtained from RHEED52,73 provided in Table 3.In Figure 3-37, parts a and b, we show the

azimuthal dependence of the (011) and (012) diffrac-tion peak intensities. The film/substrate azimuthalorientation is defined by the alignment of the (012)diffraction peak along the azimuth containing the Ausurface diffraction at φ ) 0 (see Figure 3-37b). TheBragg spacing of the (012) plane is incommensurateby 2% with respect to the nearest multiple of thesubstrate lattice spacing.

Table 5. GIXD Diffraction Peak Positions and Widths (From Ref 73)

(h,k,l) Q| (Å-1) ∆Q| (Å-1) Qz (Å-1) φ (deg)

(0,1,2) 0.8230 ( 0.0005 0.015 ( 0.0006 0.367 ( 0.003 0.05 ( 0.06(0,1,2h)a 0.8213 ( 0.0004 0.0115 ( 0.0006 0.372 ( 0.004 17.19 ( 0.05(0,2,4) 1.641 ( 0.001 0.0186 ( 0.0017 0.74b 0.3 ( 0.4(0,1,1h)a 0.6064 ( 0.0005 0.012 ( 0.001 0.2087 ( 0.005 23.39 ( 0.07(0,1,1) 0.6065 ( 0.0005 0.0139 ( 0.001 0.209b 19.64 ( 0.06(1,0,2) 0 ( 0.003 1.949 ( 0.0006

a These diffraction peaks come from different sublattices, whose orientation differs only by multiple of 60° with respect to thesubstrate. b The exact Qz position of these peaks was not measured.

Figure 3-37. GIXD data vs momentum transfer parallelto the substrate, Q|, and the azimuthal orientation, φ.Azimuthal scan of the (a) (011) diffraction peak and (b)(012) peak. (c) Radial scan of the (011) diffraction peak forφ ) 23.4° (closed circles) and φ ) 19.6° (open circles). (d)Radial scan of the (012) diffraction peak for φ ) 0° (closedcircles) and φ ) 17.2° (open circles). (e and f) : High-resolution radial scans through the (012) and (024) peaks,respectively (φ ) 17.2°). The solid lines in c-f are fits tothe data (using a Gaussian and a linear background), whilethe lines in a and b simply connect the data points. Theresolution was ∆Q| ) 0.009 Å-1 and ∆Qz ) 0.18 Å-1. In a,c, and f, the intensities are multiplied by a factor of 2 (fromref 73).

1826 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 35: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

A PTCDA film having a particular orientation(with its symmetry equivalent orientations separatedby 60°) exhibits both (012) and (012h) Bragg peaks,which have an azimuthal separation (∆φ) determinedby the lattice spacings (a,b), of the PTCDA unit cellgiven by111

where a* and b* are the reciprocal lattice vectorscorresponding to the (010) and (001) Bragg peaks,respectively. For an unstrained PTCDA lattice, a/b) 0.6007 (see Table 2) corresponding to a peakseparation of ∆φ ) 100.5°. Given the 6-fold sym-metry of the top Au layer, this results in an azi-muthal separation of 19.5°. Therefore, an unstrainedPTCDA film aligned along φ ) 0° would exhibit (012)Bragg peaks at φ ) 0°, and (012h) Bragg peaks at φ )19.5° and 40.5° (as well as at symmetry equivalentorientations separated by 60°), as observed in Figure3-37. Consequently, while the (012) and (012h) Braggpeaks are equivalent with respect to the film, theyhave inequivalent orientations with respect to the Ausubstrate (see Figure 3-38).Note that some samples show a larger number of

peaks than in Figure 3-37. For example, in somecases,111 ∆φ ) 15.2 ( 0.3°, which is significantlysmaller than ∆φ) 19.5° observed in Figure 3-37. Thenumber of inequivalent film orientations apparentlydepends upon the detailed characteristics of the Ausubstrate and growth conditions.73,111 Some of thevariability between Au substrates may be due to

variations in their surfaces, such as the magnitudeand direction of the surface miscut.From eq 3.13, the aspect ratio (a/b) of the PTCDA

unit cell can be determined. For example, ∆φ ) 15.2( 0.3° corresponds to an aspect ratio of a/b ) 0.649( 0.003. Defining strain as δ ) ([a/b]film - [a/b]bulk)/([a/b]bulk), then δ ) 8%. In the context of inorganicgrowth, this is an extreme degree of strain. Simi-larly, the aspect ratio for the φ ) 30° domain is foundto be a/b ) 0.615 ( 0.003, corresponding to a strainof δ ) 2.4%. The dependence of the lattice strainupon the azimuthal orientation of incommensuratedomains with respect to the same substrate is ademonstration of the QE nature of this film system.The X-ray diffraction intensity near the (102)

PTCDA Bragg peak as a function of the momentumtransfer perpendicular to the surface, Qz, is shownin Figure 3-39. The interplanar stacking distanceinferred from the diffraction peak position is d102 )3.22 ( 0.01 Å, consistent with bulk values for thismolecule. A rocking scan through this peak (inset,Figure 3-39) indicates that the (102) direction isaligned with the Au(111) surface normal to within<0.1°, implying that the PTCDA molecules lie flaton the Au(111) surface. The existence of fringes oneither side of the PTCDA Bragg peak implies thatthe film has a well-defined crystalline thickness overthe ∼1 mm2 area of the X-ray beam spot, while therapid attenuation of these fringes suggests somedegree of film roughness. In contrast, films grownat higher substrate temperatures do not show theseoscillations, due to complete 3D islanding. Both ofthese features will be discussed further below.While the (102) peak width, ∆Q, determines the

average thickness, L, of the films (∆Q ) 2π/L), theshape of this diffraction peak (in particular, near theshoulders) provides a measure of the disorder in thefilm thickness. These data are described by anincoherent superposition of the scattering from auniform, 17 ML PTCDA film and the Au surface,although the lack of complete destructive interference

Figure 3-38. Schematic of the (a) reciprocal, and (b) realspace, incommensurate structure of the PTCDA film onAu(111), with the unit cell noted in each. In a, the open(closed) spots indicate the positions of the PTCDA (Ausurface) diffraction conditions as a function of Q|. Thedashed and solid lines in a denote the azimuths in whichthe (0 1 (1) and (0 1 (2) peaks are found, respectively.Note that the azimuth containing the (012) diffraction peakis aligned with the Au surface diffraction peak (solid spot).The vertical solid lines in b denote the (012) Bragg planes(from ref 73).

Figure 3-39. Radial scan data as a function of Qz of the(102) Bragg peak corresponding to the lamella spacing. Theslowly increasing “background” is due to the Au surfacetruncation rod. The inset shows the rocking curve of the(102) peak, showing alignment with the surface normal ofthe Au substrate. Resolution was ∆Q| ) 0.09 Å-1, and ∆Qz) 0.009 Å-1 (from ref 73).

∆φ ) φ01h2 - φ012h ) 2 tan-1(2b*/a*) )

2 tan-1(2a/b) (3.13)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1827

Page 36: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

suggests a limited roughness of the layers. The filmsurface has been examined111 using atomic forcemicroscopy (AFM). An average domain size of ∼500nm is observed, with layer-by-layer growth within thedomains resulting in nearly monolayer flat PTCDAsurfaces. This domain size is consistent with thedimensions of the Au substrate domains, suggestingthat nucleation at substrate domain step edges maydetermine surface morphology in strained QE sys-tems.119 The PTCDA domains are separated bynarrow valleys, indicating that strain relief is occur-ring via a Stranski-Krastonov growth mechanism.Similar results, indicating even smoother PTCDA

surfaces were found by Fuchigami et al.140 In Figure3-40 are shown the X-ray diffraction patterns ob-tained as a function of growth temperature. As inthe work of Fenter,111 the PTCDA (102) planes arefound to lie parallel to the Au(111) substrate surface,and hence only the (102) and (204) reflections arevisible in these spectra. This conclusion was verifiedby measuring the relative intensities of the out-of-plane and in-plane molecular vibrations using Ramanscattering. There are very pronounced oscillationsin the GIXD diffraction patterns as the X-ray glanc-ing angle, θ, approaches 0°. These oscillations, onlyapparent for samples grown at Tsub < 100 °C, indicatethe existence of very flat organic film surfaces whengrown under nonequilibrium (i.e., QE) conditions.This improved order and surface flatness is consis-tent with electron micrographs obtained for thesesame samples.140 This study showed that the totalaverage molecular orientation improved considerablyas substrate temperature was reduced.The evolution of growth was quantified using the

intensity dependence of the specular X-ray reflectionas a function of film thickness.74 Figure 3-41 showsthe specular beam intensity obtained during PTCDAgrowth at a rate of 4.2 ML/min, and a substratetemperature of 75 °C. These intensity oscillationsresult from interference between the PTCDA and AuBragg truncation rods,161 and coincide with the onsetof film growth on the Au(111) surface. The oscilla-tions are present for 3-4 ML, after which they cease,giving way to an average intensity due to a constantlevel of coherence between X-rays reflected from thesubstrate and film surfaces. This suggests the pres-ence of a ∼3 ML thick “wetting layer” occurs underthis elevated substrate temperature, followed by

further growth of islands which eventually occupymost of the substrate surface area. As the filmthickness increases, however, the regions betweenthe layers fill in, resulting in relatively flat filmsurfaces.In contrast to PTCDA on Au(111), studies of thick

(0.4 µm) PTCDA films grown under nonequilibriumconditions (i.e., high growth rates or low substratetemperatures) exhibit extremely flat surfaces withoutany evidence for islanding when grown on amorphoussubstrates.1 Furthermore, studies of ∼300 Å thickPTCDA films on oxidized silicon substrates162 or onHOPG52 as well as some reports140 of PTCDA onAu(111) did not find any evidence for island nucle-ation. This suggests that the appearance of islandsdepends critically on details of the substrate surfaceand the strength of the organic film/substrate inter-action.72To determine the mechanism for strain relaxation

on Au(111) substrates, the PTCDA lattice dimensionsas a function of depth in thicker films have also beenexamined.73 In Figure 3-42, the (012) peak positionof an 1100 Å thick film on Au(111) is plotted as afunction of the X-ray beam incident angle. SinceX-rays exhibit total external reflection, the penetra-

Figure 3-40. X-ray diffraction patterns of PTCDA filmson Au(111)-coated LaSF6 substrates deposited at differentsubstrate temperatures. The oscillations at small θ indicateconsiderable film flatness (from ref 140).

Figure 3-41. Time dependence of the X-ray specularreflectivity at the PTCDA anti-Bragg point. Oscillations inthe intensity are clearly observed after the shutter isopened at t ) 0, and the coverages corresponding to eachsuccessive monolayer are indicated. Film growth conditionsare 4.7 ML/min at Tsub ) 75 °C (from ref 74).

Figure 3-42. Plot of the peak radial position, Q, of the(012) Bragg peak for PTCDA on Au(111) as a function ofthe incident angle, R. Also shown is the inverse penetrationdepth, 1/Λ (1/Å), as a function of R. The shift in the Braggpeak position appears when the incident angle is decreasedbelow the critical angle, Rc (i.e.,, when the penetrationdepth becomes smaller than the film thickness) (from ref73).

1828 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 37: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

tion depth, λ ≈ 40 Å, of X-rays is greatly reduced atincident angles below the critical angle of Rc ) 0.16°(for PTCDA at λ ) 1.285 Å), and hence varying thisangle provides the depth dependence of the filmlattice parameters. These data show a small mono-tonic shift in radial peak position, Q, as a function ofincident angle, indicating some minor strain relax-ation in the top ∼40 Å of the film. However, thisrelaxation is only ∆∼0.2%, whereas in this case thefilm itself is strained by at least 3.8% from its bulk(equilibrium) structure. Hence, QE films have theunique property of growing into nonequilibriumstructures determined, to some degree, by the mis-match of the adlayer lattice to that of the substrate.These nonequilibrium structures do not relax, butrather the strain is frozen in to a structure estab-lished by the first or second ML. In effect, thesubstrate acts as a template, generating a “new”crystal structure depending on the details of thesubstrate and the strength of the substrate/organicthin film interaction.Strain, as observed for PTCDA/Au(111), cannot

build up indefinitely without generating a high misfitdislocation density. One mechanism suggested insection 3.3.1.3 for strain relief is via generation ofMDWs, which result in periodic variations in one ofthe several unit cell degrees of freedom. In incom-mensurate systems, adjacent unit cells within thefirst ML are in slightly different (but spatiallycorrelated) positions with respect to the substratelattice. Hence, the degree of perturbation within aparticular unit cell to minimize energy will also bespatially correlated to neighboring cells. This spatialcorrelation should then give rise to a periodic (butsmall) variation in the positions of atoms within thesurface net. These periodic variations are a mani-festation of MDWs in large molecular systems ascompared to monatomic inert gas atom systemswhere the lattice positions themselves are periodi-cally displaced.116

While there is as yet no conclusive evidence forsuch MDWs in QE films, Fenter and co-workers73observed that the (012) PTCDA diffraction peakposition was 0.12% larger than the (012h) peak, andthe width of (012) peak was also 30% larger than thatof the (012h) peak (see Figure 3-37c-f). The broaden-ing mechanism is not fully understood, however, the(0 1 (2) Bragg planes differ only in their orientationwith respect to the Au(111) substrate (Figure 3-37d),so these structural asymmetries may reflect theperturbation of the PTCDA lattice due to its interac-tion with the Au substrate. This broadening isconsistent with a MDW since it is strongest near the(012) direction which lies along the Au symmetryaxis, and is weakest at large angles with respect tothis alignment direction, where the (012h) and (011h)peaks are found at ∼77° and ∼83° relative to the(012) peak orientation, respectively.From the foregoing discussion, it is apparent that

QE is a kinetically controlled growth mode resultingin strained, nonequilibrium structures. To explorethe conditions leading to a transition between QE andequilibrium epitaxial structures, both the substratetemperature and deposition rate were varied, andcorrelated with the resulting PTCDA on Au(111)

structure.111 It was found that there are two well-defined growth regimes: QE is achieved under non-equilibrium conditions (NEQ) of high growth rateand/or low substrate temperature, whereas un-strained epitaxy is achieved at equilibrium (EQ) withlow growth rates and/or high substrate temperature.In Figure 3-43a, lattice strain is a plotted vs filmthickness for films grown under NEQ conditions.Although the magnitude of δ is large, there is noapparent decrease of the strain even for film thick-nesses of up to∼70 ML (>200 Å). This implies eitherthat the critical thickness of the PTCDA film is verylarge or that the kinetic barrier to relieve the strainis sufficiently high that the strain is frozen-in at roomtemperature. The lack of strain relief is consistentwith studies of PTCDA/NTCDA multilayers grownunder similar conditions. There it was found163 thatthe exciton-phonon coupling, a property expected tobe extremely sensitive to strain, did not change asthe PTCDA layer thickness was increased from 10to 500 Å (see Figure 4-15).The evolution of the film surface roughness is

shown in Figure 3-44 for films grown at both high(Tsub g 100 °C, or EQ) and low (Tsub ) 17 °C, or NEQ)substrate temperatures, with a significant differencein morphology for growth under these two conditions.While NEQ films indicate a slow, monotonic evolutionof the surface width, films grown at high temperatureexhibit a much stronger dependence of the surface

Figure 3-43. Lattice strain, defined as the fractionaldeviation of a/b from bulk, as a function of PTCDA filmthickness plotted for (a) the NEQ regime (with a depositionrate of ∼4 ML/min), and (b) the EQ regime (with the 4.4and 17.3 ML films grown at Tsub ) 170 °C, while the 1 MLfilm was grown at 100 °C at a growth rate of 0.8 ML/min).In a, the strain for the (012) domain is derived indepen-dently through the separation of both the (0 1 (2) and(0 1 (1) Bragg peaks (from ref 111).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1829

Page 38: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

width for thicknesses >2 ML. The dramatic increaseof the surface roughness (particularly under equilib-rium conditions), indicates that incomplete wettingoccurs at high temperatures. In Figure 3-43b, thestrain for this epitaxial film is also shown vs thick-ness. These data indicate that strain in films of afew monolayer thickness is similar to that found forall of the nonequilibrium (NEQ) films, although thestrain is completely relieved for EQ films thicker than10 ML. Hence, the changes in film morphology andlattice strain are both due to strain relief in the thick,epitaxial (EQ) films. The solid curve in Figure 3-44is a calculation assuming growth occurs by therandom arrival of molecules with no surface diffusion.This model somewhat overestimates the surfaceroughness in the EQ regime, where a small rough-ness of ∼4 ML (rms) is observed for 30 ML thickfilms, indicating the existence of a residual surfacediffusion at 17 °C.To understand the transition between the NEQ

and EQ regimes, in Figure 3-45, the evolution of thesurface roughness is plotted as a function of thesubstrate temperature at two different growth rates.111

This plot shows that the transition is associated withan abrupt increase in the film surface width due toformation of large 3D islands as the film evolvestowards the equilibrium, incomplete wetting mor-phology. Also, at low growth rates, the roughnessin the QE regime is indistinguishable from that foundat higher rates, although the transition temperatureboundary between EQ and NEQ regimes decreasesby ∼40 °C when the growth rate is decreased by anorder of magnitude. The transition between regimes(and hence surface morphologies) results from acritical balance between adsorption and diffusionprocesses, while the film morphology is independentof the growth rate and substrate temperature withinthe NEQ (quasi-epitaxial, or kinetic) regime.

The increased order of NEQ films is also apparentfrom the X-ray pole figure plots for the more weaklybonded system of PTCDA on glass shown in Figure3-46 for two different growth rates,65 with the sub-strate held at room temperature in both cases. Thepole figure measures the crystalline texture of thesample by plotting the intensity of the intermolecu-lar, (102) Bragg reflection as a function of polar anglewith respect to the film normal. Note that AFMmeasurements of glass substrates similar to thoseused for the experiment in Figure 3-46 indicate a verysmooth surface138 with an rms roughness of <0.3 nm.From Figure 3-46, rapid, nonequilibrium growthresults in a very high degree of order, with the (102)direction tilted at 11° from the substrate normal. Incontrast, at lower deposition rates, the film is ran-domly oriented. These data are consistent with thePTCDA/Au(111) results discussed above.

Returning to Figure 3-45, we note that films grownat NEQ do not undergo a transition to the EQmorphology when annealed for several hours at Tsub

≈ 100 °C. Therefore, the transition to 3Dmorphologyappears to be due to details of the growth dynamics,suggesting that the NEQ and EQ regimes are notsimply different film phases separated by a thermalactivation barrier. This is an extremely significantfinding since it explains the apparently contradictoryobservation that high order and flat surfaces areroutinely observed at substrates cooled to very low(-100 °C) temperatures, a value well below thatwhich is expected to supply sufficient thermal energyfor molecular diffusion to occur. In the absence ofdiffusion, only amorphous growth would normally beanticipated. However, these results suggest that aresidual transient surface mobility of the PTCDAmolecules is achieved upon adsorption when thelatent heat of molecular surface condensation isconverted to kinetic energy upon adsorption. Thatis, a 3D morphology results when the moleculardiffusion length (at elevated temperatures) is largerthan the average mesa size. In contrast, when thediffusion length is smaller than the size of a typicalmesa (at NEQ), the mesa maintains the molecularlyflat terraces observed by AFM.

These observations lead to a clear picture of thethermodynamic conditions leading to NEQ, or quasi-epitaxial growth: on one hand QE films may exhibit

Figure 3-44. Root mean square surface roughness ofPTCDA films on Au(111), for NEQ films (open circles,grown at Tsub ) 17 °C) and EQ films (closed circles, grownat Tsub g 100 °C). The solid line follows evolution of filmroughness for random deposition offset by 2 ML to takeinto account the wetting layer. The dashed line is a guidefor the eye (from ref 111).

Figure 3-45. Root mean square surface roughness ofPTCDA films on Au(111) as a function of substrate tem-perature for two different growth rates (R). Note that thecross-over temperature between the NEQ and EQ regimes,as indicated by the sharp increase in surface roughnesswith Tsub, increases substantially with growth rate (fromref 111).

1830 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 39: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

a significant lattice strain. On the other hand, thereis only a weak film/substrate interaction as evidencedby the 2D azimuthal disorder. These two observa-tions can be, at least qualitatively, reconciled bynoting that organic materials are generally much“softer” than inorganic materials. For a given distor-tion, the strain energy per layer is comparativelysmaller for organic films as compared to mismatchedinorganic films. Furthermore, this strain can bereduced by taking advantage of the many degrees offreedom with the molecular unit cell. Both observa-tions are consistent with a relatively weak film-substrate interaction.3.3.2.4. Quasi-Epitaxial Growth of PTCDA on

Semiconductor Surfaces. As discussed in the previ-ous section, very strong bonding between the sub-strate and the thin film can result in epitaxialgrowth, and ultimately in considerable disorder whenthe film thickness is increased to greater than itscritical value, dc. This transition between epitaxyand QE has recently been clearly illustrated bygrowth of PTCDA on Se-passivated, and on as-grown

GaAs(100) surfaces as studied by low-energy electrondiffraction (LEED).101 In the case of the (2×4)-c(2×8)and c(4×4) reconstructed as-grown surfaces, thedangling Ga bonds generate a high density of stateswithin the bandgap at the surface of GaAs. Thesefree orbitals form bonds to the PTCDA moleculeswhich, therefore, are not free to nucleate into anordered layer. Since in this case, φintra′′ , φinter′′,amorphous or extremely fine-grained polycrystallinefilms are grown. On the other hand, a (2×1) recon-structed Se-passivated GaAs does not have anymidgap states since all surface orbitals are filled.Hence, the PTCDA surface bonding is considerablyweaker (leading to φintra′′ . φinter′′), resulting in thegrowth of an ordered, incommensurate QE layer with4-fold azimuthal symmetry with respect to the sub-strate Se dimer layer.The experiments in which this “transition” from

amorphous to QE growth on a GaAs surface wasstudied were carried out in a dual-chamber UHVsystem consisting of a MBE chamber for the growthof GaAs, interconnected with a surface analysischamber equipped with LEED, and a PTCDA effu-sion cell. For the Se experiments, 2000 Å thick layersof undoped GaAs were first grown on (100) GaAssubstrates at Tsub ≈ 550 °C. The c(4×4) and (2×4)-c(2×8) reconstructions were obtained by annealingat 250 and 450 °C, respectively. Se passivation wasachieved by in situ deposition on the as-grown GaAssurface. The PTCDA film thickness ranged between3 and 150 Å for these studies, with Tsub kept at roomtemperature during deposition. These relativelyrapid deposition rates on room temperature or cooledsubstrates result in oriented, although strained,incommensurate PTCDA QE films.73

The LEED patterns in Figure 3-47, parts a and b,were obtained from the (2×4)-c(2×8) GaAs(100)surface and from a 50-Å PTCDA overlayer grown onthat substrate, respectively. The (2×4)-c(2×8) pat-tern exhibits streaks in the [011] direction, due to acombination of point defects, steps, and displace-ments leading to kinks in the As dimer row structure.The poorly defined LEED pattern from the PTCDAlayer grown on this surface (Figure 3-47b) indicatesthat the 2D order is very poor in the plane parallelto the (100) GaAs surface. However, the presence ofthe sharp specular spot suggests regular stacking ofthe molecules in the direction perpendicular to theGaAs surface, characteristic of PTCDA films grownunder almost all conditions.54

There are significant differences between growthon reconstructed GaAs surfaces and those passivatedusing a Se layer as clearly shown for a 50 Å PTCDAlayer grown on Se-GaAs in Figure 3-48. The LEEDpattern in Figure 3-48a displays the (2×1) symmetryof the passivated surface. Previous studies of Se-passivated GaAs surfaces164-166 indicate that thesubstrate surface is characterized by 1ML of Sedimers bound to second-layer Ga atoms, a layer ofSe atoms replacing As atoms in the third layer, and0.5 ML of Ga vacancies in the fourth layer. The Sedangling bonds (DB) are filled, eliminating stateswithin the GaAs bandgap, thus reducing the substrate/molecular bond energy.

Figure 3-46. X-ray pole figure for the (102) PTCDA Braggreflection for (a) deposition on glass at a rate of g50 Å/sand (b) deposition on glass at a rate of 2 ( 1 Å/s. For a,each contour represents at 10% change in X-ray peakintensity, whereas in b, each contour represents at 1%change (from ref 65).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1831

Page 40: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

The Se passivation of the GaAs surface leads to aremarkable improvement of the 2D order in thePTCDA thin film as compared to unpassivated sur-faces. The LEED pattern of Figure 3-48b exhibitssharp spots indicative of long-range order at thePTCDA surface, and a 2-fold symmetry with respectto the substrate. The symmetry of the PTCDApattern is identical for monolayer films (3-10 Å), thinfilms (30-200 Å), and thick films (∼500 Å), suggest-ing that the PTCDA film grows from the onset in astable structure, consistent with results for PTCDAon Au(111). Considering the large lattice mismatchbetween PTCDA and GaAs, this indicates that theinterface interaction is sufficiently weak to allow thefilm to relax to near its bulk unit cell structure, anecessary condition for QE.99 Indeed, quantitativeanalysis of the PTCDA/Se-GaAs diffraction patternsuggests a unit mesh dimension which closely cor-responds to the (102) plane of the bulk crystal. Thec axis of the cell is rotated by ∼40° with respect to(but incommensurate with) the direction of the Se

dimers, as shown in Figure 3-49. Since the Se-passivated substrate reconstruction remains observ-able by LEED following the initial deposition ofPTCDA,101 as was also found for PTCDA on Au(111),

Figure 3-47. Low-energy electron diffraction (LEED)pattern from a (a) (2×4)-c(2×8) reconstructed GaAs surface;and (b) 50 Å PTCDA film grown on top of that surface.Beam energies are 65 eV for a and 13 eV for b (from ref101).

Figure 3-48. Low-energy electron diffraction (LEED)pattern from a (a) (2×1) Se-passivated GaAs(100) surface;and (b) 50 Å PTCDA film grown on top of that surface.Beam energies are 83 eV for a and 13 eV for b (from ref101).

Figure 3-49. Proposed orientation of the PTCDA latticewith respect to the Se dimers on the GaAs surface. Thereare two orientations possible, as indicated by the solid anddashed lines (from ref 101).

1832 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 41: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the interface interaction is weaker than the inter-atomic binding energies within the GaAs crystal andits surface.The 2D order of a PTCDA layer deposited on the

c(4×4) reconstructed GaAs surface is found to beconsiderably improved from that of the layer grownon the (2×4)-c(2×8) surface.101 The symmetry of thediffraction pattern is similar to that of the layergrown on the Se-passivated surface, indicating thatthe structures of the unit cells in the two cases arealso similar. In contrast to growth on the Se-GaAssurface, the diffraction spots are elongated andsomewhat diffuse, suggesting a reduced single-crystaldomain size. Nevertheless, the observation thatlarge single-crystal PTCDA domains can be grownon such “nonpassivated” surfaces is somewhat un-expected.To understand the driving forces leading to ordered

growth on these different GaAs surface reconstruc-tions, Hirose et al.101 noted that the film-substrateinteraction is determined by coupling between mo-lecular energy levels and the substrate surfacesstates,167 as shown in Figure 3-50. The ionizationenergy of the condensed phase of PTCDA is 6.4 (0.15 eV vs 5.5 ( 0.1 eV for GaAs. Assuming thatcharge transfer at the PTCDA/GaAs interface issmall, consistent with the lack of perturbation of theas-grown GaAs LEED pattern after PTCDA deposi-tion, these energies suggest that the PTCDA HOMOlies below the GaAs valance-band maximum.101

The interactions suggested for PTCDA/Se-passiva-tion GaAs are shown in Figure 3-50a. The passivatedsubstrate surface has no significant density of statesin the band gap.166 Thus the molecule-substratecoupling must occur via states that overlap with thevalence or conduction bands. However, couplingbetween the PTCDA HOMO and the filled DB(interaction 2, Figure 3-50a) should be small sinceboth levels are doubly occupied. Furthermore, cou-pling between the HOMO and empty states in theconduction band should also be weak (interaction 1)given the large energy difference between theselevels. As on H-terminated Si(111)53,91 and Se-terminated GaAs(111),81 the high degree of structuralperfection of the substrate surface added to a weakinteraction of the molecules with the inert surfaceproduces the necessary conditions for QE. Thepassivation of the substrate eliminates the chemicallyactive sites derived from unsaturated dangling bonds.The elimination of these sites weakens the molecule-surface interaction, leading to the required QE condi-tion, viz.: φinter′′ , φintra′′. A well-defined orientationof the unit cells with respect to the substrate isobserved, and corresponds to a minimum in theinterface interaction energy.The situation for (2×4)-c(2×8) and c(4×4) recon-

structed GaAs surfaces is considerably different.While no intrinsic surface states exist deep in thegap,168,169 the density of extrinsic surface defects forthe (2×4)-c(2×8) surface is large (Figure 3-50, partsb and c). The strong interaction between PTCDAmolecules and this surface was attributed in part tocoupling between the HOMO and empty and partiallyfilled gap states (interaction 1, Figure 3-50b). Thus,the lack of 2D order in the PTCDA layer is due

to interactions of the molecule energy levels withthese extrinsic and intrinsic states associated withlarge φinter′′.The weaker c(4×4) GaAs/PTCDA interaction is

due, to some extent, to the higher structural qualityof the c(4×4) surface,101 leading to a reduction indefect-induced gap states. This weaker interactionleads to the observed improvement in film order, ascompared with the more highly defected (2×4)-c(2×8)GaAs surface.A second organic/semiconductor system studied1

was PTCDA on (100)Si substrates with a very thinsurface layer of native SiO2. The presence of theoxide results in a reduction in the substrate/film bondenergy, thereby allowing for ordered film growth tooccur. Figure 3-51a shows the X-ray diffractionpatterns for 4000 Å thick PTCDA films deposited on(100)Si wafers at Tsub ) 90 and 293 K, and Figure3-51b shows the diffraction patterns for PTCDA films

Figure 3-50. Schematic band diagrams showing theproposed interactions (indicated by the solid, numberedarrows) between PTCDA molecules and (a) the (2×1) Se-passivated, (b) the (2×4)-c(2×8), and (c) the c(4×4)GaAs(100) surfaces. The PTCDA highest occupied molec-ular orbital (HOMO) is fully occupied, and the filled (cross-hatched curved regions) and empty (unfilled curved re-gions) state distributions at the GaAs surface are alsoschematically indicated to lie within the GaAs band gap(indicated by the bold lines). Ionization energies for PTCDAand GaAs are shown in a by the dashed lines. The n-typeGaAs Fermi energy, Ef, is also shown (from ref 101).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1833

Page 42: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

deposited under the same conditions on glass sub-strates. The background intensity observed in Figure3-51b is due to X-ray scattering from the glasssubstrates. For the samples deposited at 293 K, inaddition to the prominent (102) diffraction peak, asmall (110) peak is also present, indicating thepolycrystalline nature of the films. The (110) peakis observed in all samples deposited at room temper-ature or above, regardless of the substrate materialsused. On the other hand, only one diffraction peakis observed for films deposited on substrates held at90 K, indicative of significantly improved stackingorder of the PTCDA films. Using either Si or glasssubstrates, the (102) peak intensity of the samplesdeposited at low temperature is about three timesthat of the samples deposited at room temperature.The strong diffraction intensity provides evidence forimproved structural ordering in the samples depos-ited at low temperature.As in the case of PTCDA on Au(111), residual

strain is developed in organic films deposited at lowtemperatures, as observed by diffraction peak broad-ening. The full width at half maximum of the (102)diffraction peak for the samples deposited at lowtemperature is about twice that of the samplesdeposited at room temperature. Also, there is a slightpeak shift in the sample deposited on the Si substratecorresponding to strain-induced increase in the in-termolecular planar spacing of 0.026 Å. Such a shiftis not present in the sample deposited on glasssubstrates.

When films are deposited at low substrate temper-ature (Tsub < 120 K) or high deposition rate (>50 Å/s),it is apparent that molecules form stacks character-ized by long-range order, with disorder resulting fromdeposition at elevated temperatures or low depositionrates. This ordering can also be inferred from theelectrical and dielectric properties of organic-inor-ganic heterojunctions1,37,38,65 (OI-HJ). For example,it has been shown that the current (I) vs voltage (V)characteristics of OI-HJ devices at low current den-sity (J) can be modeled using:37,65,170,171

which describes thermionic emission current over thebarrier at the OI interface. Here, Vd is the voltagedrop across the heterointerface and is given by Vd )Va - Vo where Va is the applied voltage and Vo is thevoltage drop across the organic thin film. At lowforward bias where J is small, then Vd ≈ Va. Also, qis the electronic charge, and n ≈ 2 is the idealityfactor. As has been shown previously (also, seesection 5.1), Jso is the saturation current which is afunction of the OI-HJ energy barrier height.171At high forward currents, carrier transport is

dominated by space-charge effects in the organic film,in which case Vo ) Va - Vd, and

where µ is the majority carrier mobility in the organicthin film, ε is the permittivity, and d is the filmthickness. Hence, by measuring J(V) at high currentdensities where J(V) ≈ V2, the steady state carriermobility can be determined. By correlating mobilitywith film growth conditions, the degree of stackingorder can be inferred since µ is ultimately determinedby the π-bond overlap between molecules in the stack.A typical PTCDA/Si OI diode structure is shown

in the inset of Figure 3-52. Fabrication of such teststructures has been described previously.37,63 Figure3-52 shows the forward biased I-V characteristicsfor OI-HJ devices with 2000 Å thick PTCDA filmsdeposited at different Tsub. For the sample with Tsub) 90 K, the forward current has an exponentialdependence on voltage due to charge diffusion acrossthe OI heterointerface in the low bias voltage regimeand then exhibits a roll off in current due to space-charge-limited current.171 For the sample with PTC-DA deposited at Tsub ) 443 K, the forward currenthas no exponential dependence for the entire voltagerange studied, indicating a low carrier mobility. Fitsto the I-V data using analyses of the dark currentdiscussed above are also shown by the solid lines inFigure 3-52. From these fits, the carrier mobility inthe organic films decreases from 0.16 cm2/(V s) forthe sample deposited at 90 K to 2 × 10-4 cm2/(V s)for the sample deposited at 443 K, indicating acorresponding reduction in stacking order. Note thatthe highest film mobilities obtained for PTCDA are∼1 cm2/(V s) measured65 for PTCDA deposited onSi(100) substrates held at room temperature at a rateof >10 Å/s (see Figure 2-3a)scorresponding to non-equilibrium conditions expected to result in QEordering.111The mobility measurements have been explained

in terms of the thin-film structural properties. For

Figure 3-51. X-ray diffraction patterns for PTCDA de-posited (a) on Si and (b) on glass. The substrate tempera-ture used during growth is indicated in the figure. Spectrain each figure displaced along the vertical axis for clarity.The strong central peak corresponds to the (102) reflectionin PTCDA (from ref 1).

J ) Jsoexp(-qVd/nkT) - 1 (3.14)

J ) (9/8)εµVo2/d3 (3.15)

1834 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 43: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

PTCDA films deposited on glass or SiO2 at lowtemperatures or high deposition rates,65 moleculesform stacks tilted at an angle of 11° from thesubstrate normal. Due to the large π-orbital overlapalong the stacking axis, charge (primarily holes37,171-174

for PTCDA) injected into the organic films aretransported along the stacking direction, giving riseto the high mobilities observed. In contrast, filmsdeposited at elevated temperatures exhibit randomlyoriented molecular stacks. This disorder results ina reduction in orbital overlap between moleculesstacked normal to the substrate plane, therebyreducing the carrier mobility in this direction byalmost 4 orders of magnitude in some cases. Whileinferences about the stacking order can be made fromthese mobility measurements, they provide no infor-mation about the order within the plane of the filmwhere the mobility is extremely small due to negli-gible in-plane π-system overlap between molecules.In addition to the substrate temperature depen-

dence on the carrier mobility in organic films, themobility also has a thickness dependence for filmsdeposited at elevated temperatures (293 and 443 K).1Except for the films deposited at 90 K, as thethickness of the film increases, the mobility de-creases. Such a thickness dependence is also appar-ent in the surface morphologies of PTCDA filmsdeposited at room temperature.1 For example, for a500 Å thick sample deposited at room temperature,no surface texture is observed, indicating structuralordering which accounts for the high carrier mobility.On the other hand, for a 2000 Å thick film depositedat room temperature, considerable surface texture isobserved, indicating the polycrystalline nature of thefilms which give rise to a reduction in carrier mobil-ity.One consequence of the anisotropic crystal struc-

ture in QE PTCDA films is that the films exhibit

enormous anisotropies in both their conductive anddielectric properties. For example, the in-planeconductivity of PTCDA films is found to be at least 6orders of magnitude lower than the conductivityperpendicular to the film plane. It is has also beenfound that the dielectric properties of highly orderedfilms have a considerable degree of anisotropy alongdifferent crystalline directions.140,175 Provided thatthere is crystalline alignment throughout the QEfilm, the anisotropy in the dielectric constant (ε)between directions perpendicular and parallel to thethin film plane (and hence approximately perpen-dicular and parallel to the molecular stacking axis)follows: (ε⊥ - 1)/(ε| - 1) ≈ d2/L2 ≈ 0.22 due toanistropies in the dipole oscillator strength. Here,d ) 3.21 Å is the intermolecular planar stackingdistance, and L is the length of the π system alongthe perylene molecular core. This approximationassumes that m* is isotropic (which has been foundto be reasonable176,177) and that the dipole momentis due to a completely delocallized charge distributionin the extended π system. Experimentally, it hasbeen found175 that (ε⊥ - 1)/(ε| - 1) ) 0.26 for PTCDA,which is close to the value predicted by the abovemodel.Furthermore, the index of refraction measured175

at a wavelength of λ ) 1.064 µm in the directionperpendicular to the substrate plane is n⊥ ) 1.36 (0.01, whereas parallel to the plane, n| ) 2.017 (0.005, resulting in an index difference (or birefrin-gence) of ∆n ) 0.66. This structural order is main-tained along waveguides approaching 1 cm in length,as determined by a lack of rotation of the opticalpolarization vector propagating in such guides grownon amorphous Teflon22 or polyimide buffer layers.102Such remarkable long-range alignment is possiblyinitiated by ordered nucleation of PTCDA moleculesat the edges of the buffer layer strips, followed bygrowth from this initial crystalline “template”. Toour knowledge, these are the largest index anisotro-pies ever measured for thin films, giving furtherevidence for the nearly perfect crystalline orderachieved in QE growth.Further evidence for crystalline order of OMBD-

grown PTCDA films has been obtained throughmeasuring optical propagation loss in PTCDAwaveguides. In recent experiments, waveguidesdeposited on polyimide buffer layers102 (with nbuffer) 1.6 > n⊥) have losses <1 dB/cm for transverseelectric (TE) mode-polarized light, although ex-tremely high losses were measured for transversemagnetic (TM) mode propagation. This indicatesthat there is a low density of grain boundaries whichwould result in high scattering losses. Scatteringfrom grain boundaries, which is commonly observedin polycrystalline waveguides has two effects: (i)outcoupling from waveguide surfaces, and (ii) scat-tering of TE into TM modes which are extremelylossy in waveguides where nTM = n⊥ < nbuffer. Whilelow loss provides only inferential evidence for thehigh degree of crystalline order of the films, itnevertheless suggests that the number of grainboundaries in long (∼1 cm) waveguides is very small,and that the optic axis is oriented such that itsprojection is along the waveguiding direction (sinceother orientations would lead to rotation of the TE

Figure 3-52. Forward-biased current-voltage character-istics for In/2000 Å PTCDA/p-Si devices (with a top contactarea of 5.5 × 10-4 cm2) where the organic layer wasdeposited at different temperatures indicated. Solid linesare fits to the data using theory described in text. Inset:Structure of a typical OI-HJ device (from ref 63).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1835

Page 44: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

mode, and hence large optical loss). Since the opticaxis has a fixed orientation with respect to the unitcell, the molecules with the thin-film waveguide mustall be approximately oriented in the same directionrelative to each other and the buffer layer, suggestinga uniform, QE growth of the organic thin film.3.3.2.5. Other QE Systems. Several other systems

beyond the well-studied PTCDA have been reportedwhich exhibit the characteristic, incommensurate butordered growth of an organic vdW solid on aninorganic substrate typical of QE. An interestingexample system demonstrating QE growth isthat of C60 and C70 on various substrates includ-ing MoS2(0001), GaSe(0001) and H-passivatedSi(111).105,130,178 For these experiments, the filmswere studied by RHEED after growth at a rate of∼0.02 ML/s with the substrate temperature at 80 to150 °C. It was found that the layers preferentiallyalign along either the [112h0] or [101h0] substrate axes,even though the lattice mismatch ranged from 1.3%to 12.7%. Here, the hcp lattice constant of C60 is aF) 1.00 ( 0.02 nm, and for C70, aF ) 1.06 ( 0.04 nm.These results105 are summarized in Table 6. In onecase (C60 on GaSe), the mismatch for film alignmentalong the [112h0] direction was 12.7%, although C60is nearly commensurate with the [145h0] directionwhere the lattice mismatch is only 0.7%. Thisazimuthal orientation is preferred due to the lowersurface energy of the incommensurate structure.Sakurai et al.105 propose that the alignment isultimately determined by a competition betweencommensurability and the alignment along stepedges which tends to follow a principal crystallineaxis. However, it has been found72,99 that com-mensurability does not necessarily lead to the lowestenergy configuration in vdW-bonded thin films (seesection 3.3.1), since more energy is expended indistorting the film lattice than is gained by forminga commensurate lattice. This is the underlyingprinciple governing QE growth, and is a direct resultof the competition between the relatively weak,attractive vdW bond between substrate and adlayer,and the short range but strong core-core repulsionwithin the adlayer itself.Nucleation along step edges or other substrate

inhomogeneities has been observed in several experi-ments to initiate ordered QE growth independent ofcommensurability of the film with the substrate.119For example, PTCDA has been shown to align withits (100) axis parallel to the edges of polymerstrips,22,102 and perhaps even to glass.37 Nucleationfrom edges and vicinal planes on semiconductorsurfaces provides evidence that the internal energyof the thin film can, in some cases, play a greaterrole than anchoring to the substrate surface inultimately determining molecular alignment of QEfilms. This phenomenon has also been observed for

perfluorotetracosane [n-CF3(CF2)22CF3] and poly(tet-rafluoroethylene) [(CF2)n, or PTFE] on PTFE sub-strates which were first prepared by mechanicallyrubbing along one direction.179 These phenomena arenot unlike that giving rise to anchoring and long-range ordering of liquid crystal molecules on sub-strates such as polymers and glass.76

It has also been found105 that unstrained C60 filmsas thick as 100 ML grow layer-by-layer on MoS2, asdetermined from uniform, continuous RHEED pat-terns.130 While it has been suggested that C70 onMoS2 films are epitaxial due to the relatively smalllattice mismatch (1.3%), we note that such mis-matches are extremely large for more strongly bondedepitaxial layers such as group III-V semiconductors.In the absence of any evidence for strong bondingbetween C70 and MoS2 and given the inaccuraciesinherent in RHEED measurements, this apparentlycommensurate growth is not yet firmly established,and more probably is similar to the QE growth of C60on this same substrate material.As noted above, the weak interaction of PTCDA on

Au(111) leaves the surface reconstruction of Au(111)unchanged after growth.73 In contrast, deposition ofC60 on Au(111) can result in significant substratesurface reconstruction,106,180,181 suggesting that theC60 is chemisorbed, rather than vdW-bonded to thissubstrate. For example, Fartash106 has shown thatunder nonequilibrium growth conditions (i.e., atsubstrate temperatures ranging from 130 to 290 °C,and at deposition rates of ∼3 Å/s, or approximately0.5 ML/s) several different, nearly equal energyphases of C60 can coexist on the Au(111) surface.Using GIXD it was found that a commensurate phaserotated 30° relative to the Au(111) symmetry axis(R30.0) coexists with an incommensurate, in-phase(R0.0) form of C60. A third phase is also observedwith R13.0 shifting to R14.2 as the temperature isincreased above∼150 °C. The strong substrate-filminteraction results in a lifting of the Au(111) 23×x3reconstruction. This case does not, therefore, cor-respond to either QE or vdWE, since vdW forces havebeen replaced with stronger molecule/substrate bond-ing. Since the resulting film structure can be sub-stantially different than its equilibrium crystal habit,subsequent layers tend to be highly disordered,similar to the case of vdWE of Pc’s on alkali halidesubstrates.In some instances of QE growth (e.g., PTCDA on

GaAs), surface bonding can be very strong as deter-mined by thermal desorption measurements wherethe last ML of PTCDA is extremely difficult toremove except at very high substrate temperatures.54In this case, disordered growth of subsequent layersbeyond the first ML is observed. However, the firstchemisorbed ML can also serve as a “wetting” layerto the substrate, followed by subsequent, vdW-bonded

Table 6. Structure of Fullerenes on MoS2(0001) and GaSe(0001) (From Ref 105)

MoS2 GaSe

experiment closest match experiment closest match

molecule axis ∆a/a (%) axis ∆a/a (%) axis ∆a/a (%) axis ∆a/a (%)

C60 [112h0] 5.2 [112h0] 5.2 [112h0] 12.7 [145h0] 0.7C70 [101h0] 1.3 [101h0] 1.3 [112h0] 4.4 [112h0] 4.4

1836 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 45: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

QE growth of the thin film. While no strong evidencehas yet been presented for the existence of transitionregions leading to ordered QE growth, their presencecan be inferred in cases where unusually strongbinding of the adlayers to the substrate has beendetermined by such methods as thermal desorptionspectroscopy.Finally, we note that strained QE growth has also

recently been observed in Pb-stearate and Mn arachi-date thin films on mica prepared by Langmuir-Blodgett (LB) techniques.182 These films form in-commensurate, strained lattices on mica substrates,with the lattice gradually relaxing to its bulk dimen-sions as the film thickness is increased to severalmonolayers. As in the case of OMBD-grown QEfilms, the LB films are azimuthally ordered withrespect to the substrate lattice even in the absenceof lattice matching. Furthermore, these films achievetheir long-range order since the interlayer inter-actions are stronger than the substrate/filminteractionssa condition which is identical to thatleading to QE via vacuum deposition techniques. Thisobservation of QE achieved by LB deposition suggeststhat QE, and its analogous growth mode of vdWE,are general properties common to all vdW-bondedsystems.

3.4. Growth of Other Multilayer Structures byOMBD

In section 3.3.2, several examples were providedwhere alternating, ultrathin layers of two differentorganic molecular crystals were grown to form mul-tilayer stacks or structures. Since these structuresexhibit unique optical and electronic properties whichdiffer from those of bulk thin films (see section 4),we discuss here the growth of multilayer structuresbased on materials other than PTCDA and relatedcompounds considered above. There are two separateclasses of such multilayers which have been thesubject of considerable recent work: organic/organic(or heterogeneous) multilayers consisting of thecombination of layers consisting of two or moreorganic compounds, and organic/inorganic (or hybrid)multilayers which consist of an OMC and an inor-ganic molecular layer placed in an alternating ge-ometry.

3.4.1. Organic/Organic Multilayer Structures

Several phthalocyanine-based multilayer struc-tures have been studied recently.143 In that work,alternating multilayers of CuPc and MgPc weregrown by OMBD on very flat Si(100) and Si(111)substrates cooled to Tsub < 173 K, and at a rate of 5Å/min to ensure layer-by-layer growth and uniforminterface properties. The thickness of each layerranged from 30 to 100 Å. Compositional character-ization of the layers was accomplished using second-ary ion mass spectroscopy (SIMS) which could easilytrack the Cu and Mg concentrations as a function offilm depth. Clear oscillations in the SIMS profileindicated good compositional uniformity of even thethinnest layers, suggesting a lack of intermixing ofcompounds across the interfaces. High-resolutioncross-sectional TEM also indicated flat and uniform

interfaces, as is clearly shown for the 50 Å/layer ofthe CuPc/MgPc multilayer in the micrograph inFigure 3-53. Further structural analysis of thesemultilayers was performed using Fourier transforminfrared (FTIR) spectroscopy to determine the mo-lecular tilt angle with respect to the substrate plane.Here, FTIR probes molecular bonds which lie parallelto the incident light polarization vector. By tuningthe wavelength to the phthalocyanine out-of-planeC-H bond bending modes at 730 cm-1, it was foundthat the molecules are tilted at 22° with respect tothe substrate surface. That is, the Pc molecules arefound to lie at a steep inclination with respect to thesubstrate, which has also been found for PTCDAmolecules which tilt at approximately 11° withrespect to glass37 or polymer surfaces.22,102

This same group138,139 has investigated anothersystem of planar molecules based on alternatinglayers of 5,10,15,20-tetraphenylporphyrin (H2TPP;C44H30N4) and (5,10,15,20-tetraphenylporphyrina-to)zinc (ZnTPP, C44H30N4Zn) grown at a rate of 1Å/min on Si(100) substrates held at different tem-peratures ranging from 233 K to 363 K. As in thecase of perylene-based thin films, the best surfacemorphologies and order of the H2TPP films werefound at the lowest temperatures (Tsub ) 233 K),where an rms roughness of 2.2 Å, equal to that ofthe substrate, was obtained. At the highest growthtemperatures (Tsub ) 363 K), the surface morphologywas discontinuous, becoming continuous but quiterough at intermediate (Tsub ) 323 K) temperatures.Multilayer periods of 23.7 and 121 Å were grown,with the thinnest layers corresponding to the diam-eter of a single H2TPP molecule. As in the case ofPc-based multilayers,143 FTIR spectroscopy was em-ployed to check the angle of the deposited H2TPPmolecules. It was confirmed that the C-H bond tiltcould be used to infer a tilt of 75° of the molecularplane to the substrate.These authors provide excellent examples of layer-

by-layer QE growth on both Si(100) and glass (withan rms roughness measured by AFM of 2.4 Å), bothof which have previously been found65 to be suitablesubstrates on which to achieve ordered QE growth.Figure 3-54 shows small-angle X-ray diffraction(SAXD) patterns of several H2TPP/ZnTPP on Si(100)multilayer samples with different layer thicknessesand numbers of periods (ranging from 20 to 40). The

Figure 3-53. Cross-sectional transmission electron mi-crograph of a multilayer structure consisting of 50 Å CuPcalternating with 50 Å MgPc layers deposited on a Si(100)substrate. The contrast between the MgPc and CuPc layersis apparent (from ref 143).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1837

Page 46: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

substrate was dipped in 1% HF in H2O solutionimmediately prior to growth to remove the nativeoxide. Similarly, Figure 3-55 shows a SAXD patternfor a H2TPP(11 Å)/ZnTPP(110 Å) multilayer grownon glass. For both substrates, the clearly resolvedoscillations at small X-ray incident angles indicatesextremely good coherence between the X-rays scat-tered from the various layers due to very flat anduniform interfaces over the entire X-ray beam cross-sectional area. These oscillations, in conjunctionwith SIMS data138 indicating abrupt compositionaldiscontinuities between each layer (i.e., no evidencefor molecular intermixing between layers), suggestthat this system does not require lattice matchingto achieve molecularly flat interfaces with long-rangeorder, as expected for QE growth.Akimichi et al.183 investigated the growth of an-

other molecular system consisting of pentacene/tetracene multilayers on glass substrates. Here, thegrowth rate was 1.6 Å/s for the tetracene layers, and

0.2 Å/s for pentacene on glass held at room temper-ature. Multilayers consisted of layer thicknessesranging from 14 ML to <1 ML, and the number ofperiods ranged from 10 to 200. Both X-ray diffractionand photoluminescence spectroscopy were used toinvestigate the resulting thin-film structures. TheX-ray spectra were obtained in the Bragg-Brentanogeometry (i.e., θ - 2θ scans), and it was found thatthe diffraction peak due to intermolecular stackingwas surrounded by satellite peaks when more thanthree molecular layers were grown in a given stack.These satellite peaks were attributed to the existenceof a well-defined superlattice period. Modeling of theX-ray diffraction spectra for ideal superlattices wasconsistent with this conclusion, providing evidencethat the interfaces between layers of differing mo-lecular composition were flat due to a layer-by-layergrowth mode. However, for superlattices consistingof layers less than three molecules deep, the satellitepeaks vanished due to compositional intermixing. Allof these findings were confirmed by photolumines-cence spectra of the samples, which suggest materialintermixing in the thinnest layer structures. Thisfinding sets a lower limit on the thickness of thesuperlattices that can be achieved with this materialssystem.Growth of organic multilayer structures with mono-

layer control has recently been demonstrated bytaking advantage of a combination of substratetemperature control and the differential sublimationtemperatures of various molecular species employedin the superlattice.50,184 The method of growth, calledmolecular layer deposition, or MLD, is analogous toatomic layer epitaxy (ALE) employed for inorganicsemiconductor MQW structures,185-188 where thegrowth of each atomic layer is self-terminating dueto the thermodynamic conditions or the bondingnature of the various molecules employed. In effect,MLD is “self-assembly” in the vapor phase. Yoshimu-ra, Tatsuura, and Sotoyama50 demonstrated themonolayer control of growth by employing pyromel-litic dianhydride (PMDA, or C10O6H2) alternated witheither 2,4-diaminonitrobenzene (DNB) or 4,4′-diami-nodiphenyl ether (DDE). These molecules werechosen due to their considerably different sublimationtemperatures (Ts), and hence their deposition iscontrolled by adjusting the substrate temperature.That is, when Tsub > Ts, the molecular stickingcoefficient is small, resulting in no deposition. Onthe other hand, by predepositing a molecule (moleculeA) whose bond energy with molecule B (EA-B )kTA-B), is higher than the bond energy of the mol-ecule to itself (EB-B ) kTB-B), and then adjusting thesubstrate temperature such that TB-B < Tsub < TA-B,a condition is created where only a single (butcomplete) monolayer of molecule B can be depositedon a layer of molecule A. That is, once a saturatedML of molecule B is deposited, no further depositioncan occur since the substrate temperature is Tsub >TB-B. To deposit a second monolayer of molecule A,the substrate temperature is adjusted to TA-A < Tsub< TB-A, and the process is repeated. All steps result,therefore, in the saturated growth of single mono-layer film stacks in an alternating, layer-by-layerprocess.

Figure 3-54. Small-angle X-ray diffraction spectra ofseveral different superlattice films grown on Si(100) sub-strates at Tsub < 233 K: (a) 20 periods of H2TPP(11 Å)/ZnTPP(33 Å); (b) 20 periods of H2TPP(33 Å)/ZnTPP(11 Å);(c) 40 periods of H2TPP(11 Å)/ZnTPP(11 Å) (from ref 138).

Figure 3-55. Small-angle X-ray diffraction spectrum of aH2TPP(22 Å)/ZnTPP(22 Å) superlattice grown on a Corning7059 glass substrate (from ref 138).

1838 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 47: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

The process of MLD of PMDA and DNB is shownschematically in Figure 3-56. Here, the carbonyl-oxy-carbonyl groups of PMDA combine with theamino group in DNB (or DDE) to form alternatelyconnected polymer chains. In contrast, the carbo-nyl-oxy-carbonyl and amino group self-bond ener-gies are relatively weak, fulfilling the major require-ment of this growth process. To accomplish thesaturated alternating monolayer growth for thismaterials system, the substrate temperature was setat Tsub ) 50 °C which results in the thermal energyexceeding that of the internal bond energy of DNB,thereby allowing for only a single ML of DNB to formon the predeposited PMDA layer (Figure 3-56a).Completion of a monolayer was clearly indicated bya crystal microbalance thickness monitor whichserved as the substrate for the organic films. Next,the substrate temperature is adjusted to Tsub ) 65°C which results in the desorption of the DNB. This

process of adsorption followed by desorption wasevidence that a single, saturated ML growth wasachieved. The alternate growth of PMDA on DNBcould be accomplished at Tsub ) 80 °C, as shown inFigure 3-56b. Successive growth of PMDA/DDEmultilayers were grown by this method, as shown inFigure 3-56c, resulting in a 100 Å thick continuouslyalternating (PMDA-DDE)n polymer (with n ) 15)oriented in the surface-normal direction. While thisprocess can be slow and difficult to control, especiallyfor molecular combinations whose sublimation ener-gies are close to the intermolecular bond energies, itnevertheless is an exciting proof-of-principle demon-stration of saturated layer-by-layer control of growthin the vapor phase. This process has the potentialfor adaptation to laser-assisted growth, where thetemperature can be locally and rapidly varied toinitiate growth which is heterogeneous in both thesurface normal and in-plane directions.

Figure 3-56. Molecular layer deposition sequences for two different polymer structures: (a) Film thickness vs shutteroperation for DNB growth on a PMDA layer (left) and PMDA growth on a DNB layer (right) and (b) polymer film growthby MLD using PMDA and DDE as the precursors (from ref 50).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1839

Page 48: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

3.4.2. Organic/Inorganic Hybrid Multilayer Structures

Many interesting physical properties of hybridorganic/inorganic multilayer structures have beenpredicted, including unique nonlinear optical proper-ties resulting from a new excitationsthe hybridFrenkel-Wannier exciton existing at the organic/inorganic interface.160,189 Furthermore, such hybridcombinations may provide means for combining theoptimal properties of very heterogeneous materialssystems in ways which, to date, have not beenanticipated. Hence, controlled growth of hybridorganic/inorganic superlattices represents an entirelynew approach to tailoring materials properties, whileproviding unprecedented challenges to the materialsscientist in achieving flat, uniform interfaces betweenmaterials with vastly different structural and chemi-cal properties.There are several issues which must be considered

in the growth of these structures. The primaryconcern is the large difference in evaporation tem-peratures of organic and inorganic materials. Forexample, Ti and Si used in CuPc/oxide multilayers190are typically evaporated at Knudsen cell tempera-tures of 2270 and 1700 K, respectively. This isconsiderably higher than the temperature (650 K)used for CuPc sublimation. The highly energeticatomic species can therefore melt, sublime, or de-compose the underlying organic layer during thedeposition process, resulting in disordered structuresor rough interfaces, thereby seriously degrading thecontrol over the properties of the structure. Further-more, many crystalline ceramics or other inorganicmaterials must be deposited on substrates main-tained at high temperatures. Once again, this hightemperature may be incompatible with the physicalcharacteristics of the organic materials used, result-ing in sublimation from the substrate, or evendecomposition. Finally, incompatibilities in materi-als chemistry can result in unacceptable or unantici-pated changes in the properties of the materialscomposing the hybrid structure. For example, CuPcis known to be doped by oxygen.21 In the fabricationof oxide/CuPc multilayers, therefore, we anticipatethe electrical properties of the CuPc will be stronglydetermined by the details of the deposition process.In spite of these difficulties, successful growth of

hybrid organic/inorganic superlattices has alreadybeen achieved. Perhaps the most interesting dem-onstration has been the growth of CuPc/TiOx andCuPc/SiO2 multilayer structures.144,190,191 Here, thedeposition of the metals was done in an O2 ambientmaintained at 2× 10-4 Torr, with the deposition rateof all constituents at 1 Å/s. To ensure that only theR form of CuPc was grown, the Si substrate temper-ature was Tsub ) 200 °C. This low substrate tem-perature resulted in amorphous oxide growth. In thecase of TiOx (with x ≈ 1.5 for these conditions),crystalline growth occurs at Tsub > 400 °C, resultingin an electron mobility of 1-2 cm2/(V s). In contrast,the mobilities of the amorphous materials grown byTakada and co-workers were too low to be deter-mined. Twenty period multilayer stacks consistingof alternating layers of 50 Å CuPc followed by anequal thickness of either the Ti or Si oxide weregrown.

The structures were studied by SIMS, X-ray dif-fraction, AFM, and TEM. The AFM images of themultilayer stacks indicated very flat interfaces witha roughness of <7 Å.144 However, asymmetriesbetween the interfaces were found using this imagingtechnique. The TiOx/CuPc interface was not as easilyscratched by the AFM tip as was the CuPc/TiOx

interface, suggesting some interpenetration (andsubsequent “hardening”) of the CuPc surface by thedeposited Ti atoms. While some interface disruptionwas inferred from this observation, the depth towhich the interface was affected was beyond theresolution of the X-ray and SIMS measurementsmade, suggesting very limited extent to the damage.Indeed, the quality of these interfaces is readilyapparent from the remarkable cross-sectional trans-mission electron micrographs of several differentCuPc/oxide samples shown in Figure 3-57. A cross-sectional view of a CuPc/SiO2 superlattice is shownin Figure 3-57a, providing direct observation of thesuperlattice structure. The layers appear to berelatively uniform, although the interface roughnessis readily observed for the individual, 50 Å thicklayers. A more detailed micrograph of a CuPc/TiOx

interface is shown in Figure 3-57b. Here, an inter-face roughness of approximately 1ML (as determinedover the∼500 Å range of the micrograph) is observed.In addition, the four individual CuPc molecularlayers are clearly observed in the lower half of theimage. By depositing submonolayer thicknesses ofCuPc on TiOx, island growth can be observed, asshown in the micrograph in Figure 3-57c. Theseislands are in fact small, highly ordered structureswith dimensions of 20 nm × 60 nm completely buriedin a TiOx matrix which is deposited over the CuPc.Since the organic material is surrounded on all sidesby the oxide, this appears to be a means to groworganic “nanoboxes” and other low-dimensional struc-tures once only thought possible using inorganicsemiconductor nanostructures.In a second series of investigations, heterogeneous

multilayer structures consisting of 3,4,9,10-peryle-netetracarboxylic diimide (PTCDI) or CuPc combinedwith the low dielectric constant metal fluoride, MgF2,were investigated.145 The substrate used was Corn-ing 7059 glass. Deposition of organic films rangingin thickness from 5 to 50 Å was done at a rate of 20Å/min on substrates maintained at room tempera-ture. The alternating MgF2 film was deposited at thesame rate to a thickness of 30 Å in all samples. Thestructures consisted of 20 superlattice periods. Analy-sis of the grown structures was accomplished usingSAXD and FTIR. Good layer thin uniformity wasobserved even for the thinnest PTCDI and CuPclayers as determined by small-angle Bragg reflectionsfrom the repeating superlattice structure. Further,analysis of the FTIR spectrum of PTCDI indicatedthat the molecules stack nearly parallel to the surfaceplane. In contrast, no such uniform stacking of CuPcwas observed. This was attributed the coexistenceof both the R and â forms of the CuPc crystal,21,118resulting in considerable structural disorder. Theoptical properties of these structures was also de-pendent on film thickness, as will be discussed insection 4.3.

1840 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 49: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

This group159 also investigated the growth andoptical properties of 30 period, tris(8-hydroxyquino-line)aluminum (Alq3)/30 Å (MgF2) superlattices, wherethe amorphous Alq3 layer thickness was varied from

10 to 50 Å, and deposition was on either glass orSi(100) substrates at a rate of 20 Å/min. As in thecase of the PTCDI and CuPc multilayers, thesestructures were investigated by small-angle X-raydiffraction, with spectra shown in Figure 3-58. Verygood agreement was found between the designedlayer thickness and that inferred from the positionsof the small angle Bragg peaks observed. StrongBragg reflections are observed even for the thinnest(10 Å) Alq3 layers, indicative of flat and uniforminterfaces. Here, layer nonuniformity would resultin a broadening and attenuation of the peaks whichis not seen in these diffraction patterns.

3.5. Summary of Growth ResultsUltrahigh vacuum, organic molecular beam deposi-

tion has been shown to be a highly versatile tech-nique for the growth of a wide range of organic andhybrid organic/inorganic thin film structures whichcomprise a promising new class of engineered mate-rials for optical and electronic device applications. Inthis section, we have compared the growths of severaldifferent archetype organic compounds on a varietyof substrates. We find that, depending on relativestrengths of the bond interactions within a film, orbetween the film and substrate, the resulting filmlattices are commensurate (i.e., epitaxial or vdWE),or are incommensurate (i.e., quasi-epitaxial) with thesubstrate. Epitaxial films are bonded to the surfaceby a combination of vdW and electrostatic forces,whereas vdWE film binding is primarily via only thevdW force. Finally, in spite of the incommensurabil-ity of QE films, they can nevertheless exhibit long-range order and have a unique orientational rela-tionship with the substrate lattice.With regard to QE, we have focused to a consider-

able degree on several aspects of film growth, from

Figure 3-57. Cross-sectional TEM images of severaldifferent multilayer structures: (a) CuPc/SiO2 superlatticepealed off of a Si substrate. Bright and dark regionscorrespond to the CuPc and SiO2 layers, respectively. (b)High-resolution image of a single TiOx layer on CuPc. Fourmonolayers of CuPc can clearly be resolved beneath theTiOx. (c) CuPc “nanobox” in a TiOx matrix. In the lowerview is a schematic representation of the structure and anAFM scan showing the surface profile of the buriedstructure (from ref 190).

Figure 3-58. Small-angle X-ray diffraction patterns formultilayer structures consisting of 30 Å thick MgF2 layersalternated with Alq3 layers whose thicknesses are indicated(from ref 159).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1841

Page 50: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the monolayer to the bulk thin film stage for themodel planar molecules: PTCDA and NTCDA. Struc-tural data obtained from monolayer and multilayerthin-film stacks indicate ordering into structureswhich are consistent with predictions made usingsimple energy-based models describing QE growth.The agreement between theory and measurementindicates that the fundamental assumptions regard-ing quasi-epitaxial growth (i.e., that the interlayercompressibility be much larger than the compress-ibility within a layer) are supported by observationfor large, planar, van der Waals-bonded molecularsolids such as PTCDA and NTCDA. These theoreti-cal predictions have also been extended to the growthof two dissimilar organic molecules. The experimentspresented here on multilayer stacks of PTCDA-on-NTCDA also provide evidence to support the theo-retical results. The high degree of ordering stronglyinfluences such macroscopic film properties as opticalbirefringence, a pronounced anisotropy in the dielec-tric and conductive properties of the films, filmmorphology, and carrier mobility.Taken in the context of a growing body of evidence

for QE growth of a large number of molecular specieson a wide variety of substrates, it is apparent thatQE growth of van der Waals-bonded planar organicmolecules is a general property of these materialssystems. Since this property has been shown tostrongly influence the macroscopic optoelectronic filmproperties (see section 4), quasi-epitaxially grownorganic thin films may open the way to an unprec-edented range of applications in photonics and elec-tronics. Strain energy in QE films is relieved bysubtle variations in the internal lattice degrees offreedom. While QE films are nonequilibrium struc-tures which differ from the bulk structure of theOMC, they are nevertheless found to be stable overa wide range of conditions such as temperature,pressure, etc. Hence, QE may result in an entirelynew class of materials whose structural, and henceoptical and electronic properties may be tuned simplyby the choice of substrate or growth conditionsemployed.In some respects, QE growth is similar to “orien-

tational epitaxy” observed for inert gas atoms phy-sisorbed onto graphite surfaces, and also is similarto the ordering of liquid crystal molecules to glassand polymer surfaces. By studying the commonfeatures of these seemingly disparate growth phe-nomena, we can gain a deeper understanding of theunderlying physics, and hence improve our controlover vdW film growth.

4.0. Optoelectronic Properties of OMBD-GrownOrganic FilmsHaving demonstrated that OMBD is a useful tool

to grow well-defined structures with interfaces be-tween layers which are flat on the molecular scale,the next step is to determine how the optoelectronicproperties of such organic nanostructures differ fromthose of bulk films. In particular, as the film thick-ness approaches that of an electron de Brogliewavelength, we anticipate that the films will exhibitquantum phenomena which are not observed inthicker (or “bulk”) films. However, we note that such

quantum phenomena will be much more difficult toobserve in organic thin films than in conventionalsemiconductors or metals due to the localized natureof excitations in vdW-bonded materials such asOMCs. For this reason, the structures need to becorrespondingly smaller in OMCs before quantumsize effects can be observed and possibly exploitedfor device applications.In the previous section, we have demonstrated that

organic thin films can be controllably grown todimensions of only a few monolayers. Hence, choos-ing archetype, extremely closely packed organicsystems such as PTCDA or bis[1,2,5]thiadiazolo-p-quinobis(1,3-dithiole) (BTQBT)192-194 which mightexhibit a developed and extended band structure,provides opportunities to study, for the first time, thenature of fundamental quantum excitations in OMCsusing precision nanostructures. In this section wewill provide a detailed review of such studies ofPTCDA-based multilayers, along with conclusiveexperimental and theoretical evidence that funda-mental optical excitations (called “excitons”) canindeed be modified in certain nanostructured OMCsystems characterized by a high degree of π-orbitaloverlap between adjacent molecules. Furthermore,we will review data relating to “quantum confine-ment” in organic nanostructures and discuss the classof molecules where such phenomena are expected tobe clearly observed. We note that virtually all suchdata accumulated thus far relate to the quantumconfinement of excitons in ultrathin organic layers.However, quantum confinement of charge carriers(i.e., electrons and holes) is also possible, and shouldlead to the observation of a broad new range ofelectronic properties unique to organic nano-structures.25-27,29

4.1. Excitons in Ultrathin Organic Films Grownby OMBDThe nature of excitons in OMCs has been the

subject of intense debate for over thirty years.195 InFigure 4-1, we illustrate the three types of excitons

Figure 4-1. Schematic representation of the various typesof excitons in semiconductors and insulators. The cross-hatched circles represent molecular lattice positions, withlattice constant aL. (a) Small radius, or Frenkel exciton withradius a ≈ aL. This exciton corresponds to a molecularexcited state. (b) Large radius (a . aL), hydrogen-like,Wannier-Mott exciton commonly found in inorganic semi-conductor materials. (c) Intermediate radius (a > aL) chargetransfer (CT) exciton proposed to lead to photoconductivityin many organic molecular crystals. The “figure-eight”within each excited state represents the molecular p bond(from ref 196, p. 71).

1842 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 51: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

found in crystalline solids. Excitons in OMCs havebeen described as either Frenkel or charge-transfer(CT) states.196 A Frenkel exciton is a correlatedelectron-hole pair localized on a single molecule,whereas the electron in a CT state is correlated witha hole located at a neighboring molecular site. In thissense, CT excitons resemble Wannier excitons com-mon in uncorrelated inorganic systems, where theinteraction energy between atomic cores is large andthe radius of the exciton is often more than an orderof magnitude greater than the interatomic separa-tion. Independent of the scale, these fundamentalexcitations are called excitons since they share theproperty of having a nonnegligible mobility withinthe crystal. Due to its spatial extent, it is possiblefor the CT exciton binding energy to be modified fromits bulk value when the electron-hole pair separation(or exciton radius) is on the order of the film thick-ness. On the other hand, we do not expect sucheffects to be observed for Frenkel states due to theirstrong localization at a single molecule. Given thatOMBD can create nanostructures on the scale of theexciton radius, we now have the opportunity to studymany of the details of these fundamental excitationswhich have previously been inaccessible using bulkstructures, thereby providing a deeper understandingof the physics of carrier transport and the opticalproperties of a large class of organic molecular solids.With this perspective in mind, considerable work

on quantifying the optical properties in OMBD-grownnanostructures has been carried out over the lastseveral years. These studies have focused primarilyon highly controlled, relatively defect free, and well-characterized “model” compounds based on OMBD-grown PTCDA thin films and its analogs. In thissection, we will discuss some of these results in detailsince they lead to an entirely new perspective andunderstanding of fundamental excitations in OMCswhich have only been possible due to the precisecontrol of structure and material purity throughOMBD growth. At the end of this section, we willdiscuss phenomena in other, less thoroughly char-acterized organic-based multilayer structures.

4.1.1. Spectroscopic Identification of Excited States inPTCDA

The first step in understanding quantum sizeeffects in ultrathin, OMBD-grown organic films is tounambiguously identify the various features in themolecular exciton spectrum. For this purpose, theabsorption and fluorescence spectra of both films anddilute solutions of the model compound, PTCDA,have been thoroughly studied.197 Thin films weregrown on cooled (∼85 K) glass substrates followingprocedures discussed in section 2. Solution sampleswere prepared by dissolving prepurified PTCDAcrystalline powder in dimethyl sulfoxide, (CH3)2S:O(DMSO) and other solvents via sonication. PTCDAis found to be weakly soluble in DMSO, ensuring asmall solvent/molecule interaction, while limiting themost concentrated solution to only 2 ( 1 µM.The absorption spectrum of the 2 µM stock solution

in DMSO, shown in Figure 4-2a, has four clearlyresolved peaks with a shoulder on the high energyside of the 2.7 eV peak. In addition, the low-energytail (defined as the energy at which the absorption

is >5% of its peak value) extends to E < 1.90 eV.This is compared to the thin-film absorption spec-trum (Figure 4-2b) with a broad feature at E > 2.3eV and a distinct, narrow peak centered at E ) 2.23eV. Furthermore, the low-energy tail is compara-tively short, extending to only E ≈ 2.0 eV.There are significant differences between the fluo-

rescence spectra of PTCDA solutions and thin films.The solution fluorescence spectrum (Figure 4-2a) hasa maximum at E ) 2.32 eV, with a shoulder at E )2.18 eV. In contrast, the thin-film short wavelengthfluorescence is completely quenched, with the peakfluorescence centered at E ) 1.70 eV, and with anadditional feature at E ) 1.55 eV. No significantchanges in the shape of the thin-film or solutionfluorescence spectra were observed as the pumpenergy was changed from E ) 2.72 to 2.05 eV.The absorption was also studied as a function of

concentration of PTCDA in solution (Figure 4-3a). Toquantitatively study the spectral features, each solu-tion absorption spectrum was fit to six Gaussianpeaks with the parameters provided in Table 7. Thefour high-energy Gaussian curves correspond to thedistinct absorption peaks, whereas the fitting pa-rameters for the two lowest curves at E ) 2.23 eVand 2.12 eV were chosen to agree with electroabsorp-tion177 and photoconduction172,173 data in PTCDA thinfilms. Table 7 also shows the results of a similar fitto the thin-film absorption spectrum, where the twolowest energy peaks are obtained from these samestudies.

Figure 4-2. Absorption (solid line) and fluorescence(dashed line) spectra of (a) 2 µM solution of PTCDA inDMSO, and (b) 1000 Å thick PTCDA film. The energyposition for each absorption and fluorescence transition isidentified (from ref 197).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1843

Page 52: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

The dependence of the E ) 2.23, 2.39, 2.55, and,2.74 eV integrated peak areas (which are propor-tional to the transition oscillator strength) as afunction of PTCDA concentration [k] in DMSO areshown in Figure 4-3b. Both the E ) 2.74 and 2.23eV peaks rapidly decrease with respect to the E )2.39 eV peak, however only the 2.23 eV peak com-pletely attenuates at concentrations of [k] e 0.25 µM.For the three higher energy peaks, the rate of

decrease in absorption with decreasing PTCDA con-centration is a function of energy, where the 2.39 eVpeak shows the smallest rate of decrease.The excitation spectrum of the luminescence peak

centered at E ) 1.70 eV of a 650 nm thick film isshown in Figure 4-4. Low fluorescence efficiency(<1%) is observed at E < 2.2 eV, with a peak at ∼20times its short wavelength efficiency at 2.10 eV,which has been attributed to a bulk rather than asurface phenomenon.197The configuration space representations of the

molecular levels in PTCDA shown in Figure 4-5 areused to interpret these solution and thin-film spectra.Here, Figure 4-5a corresponds to the absorption

Figure 4-3. (a) Absorption spectrum of PTCDA in DMSOsolution as a function of PTCDA concentration. Eachdilution step reduces PTCDA concentration to 80% of thestarting solution. (b) Relative oscillator transition strength(proportional to the integrated area) of E ) 2.20 eV, E )2.39 eV, E ) 2.55 eV, and E ) 2.74 eV solution absorptionpeaks as a function of PTCDA in DMSO concentration. Theslopes refer to the linear fits (straight lines) to the data.The plot was normalized to the oscillator strength of theS1 [0-0] transition at a concentration of 2 µM (from ref 197).

Table 7. Numerical Fit to the Absorption Spectra of PTCDA Thin Films, and 2 and 0.25 µM Solutions of PTCDAin DMSOa

thin film 2 µM solution 0.25 µM solution

transition E0 (eV) ∆E (eV) FWHM (eV) f E0 (eV) ∆E (eV) FWHM (eV) f E0 (eV) ∆E (eV) FWHM (eV) f

CT [0-ST] 2.11 0.17 0.46 2.12 0.36 1.10 2.16 0.50 0.63CT [0-F] 2.23 0.14 2.03 2.20 0.16 0.73 2.20 0.16 0.06S1 [0-0] 2.38 >0.11 0.14 1.00 2.39 >0.16 0.10 1.00 2.39 >0.17 0.09 1.00S1 [0-1] 2.49 >0.17 0.21 2.86 2.55 >0.19 0.16 1.58 2.56 >0.17 0.13 1.03S1 [0-2] 2.66 >0.19 0.23 3.20 2.74 >0.16 0.16 0.94 2.73 >0.17 0.12 0.39S1 [0-3] 2.85 0.26 3.31 2.90 0.13 0.22 2.90 0.23 0.41

a Each spectrum was fit to six gaussian shapes corresponding to six transitions identified in the text and noted in the left-mostcolumn. For each gaussian curve, the peak energy (E0), energy difference of the neighboring energy peaks (∆E), the full widthat half maximum (FWHM), and the integrated area below the curve (f) normalized to the area of S1 [0-0], are indicated. (FromRef 197).

Figure 4-4. Excitation fluorescence spectrum of a 650 nmthick PTCDA film sample (from ref 197).

Figure 4-5. (a) Absorption and (b) fluorescence configu-ration space diagrams along with direct transitions pro-posed for PTCDA (from ref 197).

1844 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 53: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

transitions, whereas Figure 4-5b provides the fluo-rescence transitions occurring after nuclear relax-ation. The electronic ground state of PTCDA is spin-singlet (labeled S0 in Figure 4-5). The high oscillatorstrength of the absorption features in Figure 4-2 istherefore a signature of spin-allowed singlet-to-singlet transitions. The distinctly different behaviorof the high-energy solution absorption spectra at E) 2.39, 2.55, 2.74, and 2.90 eV, as compared to thebroad peak at 2.23 eV as a function of solutionconcentration suggests that these two groups oftransitions are associated with separate manifolds.The equal spacing between the four high-energysolution absorption peaks (c.f., Table 7) arises sincethey originate from transitions between differentvibronic states (νn) in the singlet, S1 band, labeledS1[0-νn] (with νn ) 0, 1, 2, 3). In solution, theabsorption spectral shape of these S1 transitions isnot significantly influenced by the PTCDA concentra-tion in DMSO (Figure 4-3a), indicating that S1 isFrenkel-like, i.e., the excitation is largely confinedto the individual monomers.In contrast, the absorption centered at E ) 2.23

eV is strongly affected by concentration, becomingvanishingly small for [k] < 0.1 µM solutions. This isexpected for an aggregate state arising from the π-πorbital overlap of closely stacked PTCDA molecules,and is similar to previous observations of absorptionby aggregate states in other organic crystals198-200

such as R-perylene, tetracene, and C60. Since PTCDAis only weakly soluble in DMSO, small particles withthree-dimensional crystalline properties (e.g., ex-tended CT states) form a suspension in solution. Asthe colloidal solution is diluted, the concentration ofaggregate particles is reduced to the extent that, atthe lowest concentrations, their contribution to theabsorption spectrum is negligible. Thus, the 2.23 eVabsorption feature has been assigned197 to a charge-transfer (CT) exciton of the extended PTCDA crystalstructure. The lower energy of the 2.23 eV peak ascompared to the S1 state is also consistent with CTstate properties.195

Since monomers are the dominant species in themost dilute solutions, and since the fluorescencespectrum does not significantly change with concen-tration, the solution fluorescence in Figure 4-2a hasbeen attributed to direct monomer transitions fromS1 to S0 (see Figure 4-5b). In contrast, the thin-filmfluorescence spectrum peaks at 0.60 eV lower energythan does the monomer fluorescence (Figure 4-2b).The fluorescence quantum efficiency of the thin filmshas been found to be significantly lower than that ofsolutions due to phonon-enhanced fluorescencequenching (internal conversion). Therefore, despitea small aggregate concentration in PTCDA solutions,only a very low fluorescence yield has been observedfor the aggregate peak in solution.197 The S1 absorp-tion lines of aggregates and thin films are broaderthan for monomers due to the increased coupling ofphonon modes with electronic transitions in solids.Further, since the fluorescence quantum efficiencyin the excitation spectrum peaks at 2.10 eV corre-sponding to the low-energy tail of CT (c.f., Figure 4-4),the CT state is likely the initial level in the fluores-

cence process, and there are nonradiative losses inthe S1 f CT transitions.These observations suggest that most S1 excitons

undergo either S1 f S0 internal conversion that issufficiently fast to effectively compete with fluores-cence, or intersystem-crossing into a nonradiativetriplet state, T1, which is typically lower in energythan S1. Figure 4-6 summarizes the various possibletransitions between the lowest energy states foundfor PTCDA thin films. Both internal conversion forS1 f S0 and the spin-forbidden T1 f S0 nonradiativetransitions lead to the low quantum yield of solidPTCDA.In molecular crystals such as pyrene and perylene,

strong exciton-phonon coupling results in the forma-tion of self-trapped excitons. These states are lowerin energy than the corresponding free excitons andcan be solely responsible for the observed fluores-cence. Self-trapping in PTCDA has also been identi-fied197 by analyzing the low-energy thin-film absorp-tion (or Urbach) tail of the CT state.201 Due to strongexciton-phonon coupling, self-trapping is very likely.In that analysis, a small self-trapping energy of ∆EST) 0.04 ( 0.02 eV was measured, placing the self-trapped CT exciton (denoted CTST) at 2.19 ( 0.05 eV,which is in reasonable agreement with 2.11 ( 0.04eV obtained through numerical fits to the absorptionspectrum.197 In contrast, the small radius S1 excitonis not subject to long-range lattice interactions, andhence does not self-trap. From the small Stokes shift(<0.1 eV) between the solution peak absorption andfluorescence spectra (Figure 4-2a), it is likely thatrelaxation occurs directly from the free S1 state intoS0, which can also be inferred from the short solutionfluorescence lifetime of 4.0 ( 0.5 ns characteristic ofsuch transitions.202 In contrast, the self-trapped CTstate of PTCDA thin films have a fluorescencelifetime24 for excitation into S1 approximately threetimes longer, at 10.8 ( 0.5 ns. The relatively shortlifetime of solution luminescence implies a lack ofsignificant self-trapping for S1.

4.1.2. Fluorescence and Absorption Spectra of PTCDA/NTCDA MQWs

Having unambiguously identified the CT excitonin PTCDA as an extended, or aggregate state, weexpect its energy to be changed when it is confinedin nanostructures whose dimensions are comparableto the effective exciton radius, a*. That is, bygrowing ultrathin films of dimension ∼a*, we antici-pate that the energy barriers at the film surfaces will“confine” the exciton within that film, and henceincrease its binding energy accordingly. These con-

Figure 4-6. Proposed transitions between the lowest lyingthin-film PTCDA energy levels. The straight and wavylines indicate single-step and multi-step nonradiativetransitions, respectively (from ref 197).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1845

Page 54: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

finement phenomena were explored by measuring thespectral properties of PTCDA/NTCDA and PTCBI/NTCDA multilayer structures, where the film thick-nesses were varied from 10 to 500 Å.163 Here, PTCBI(or 3,4,9,10-perylenetetracarboxylic bisbenzimida-zole) has a larger interplanar stacking distance of3.45 Å as compared with 3.21 Å for PTCDA, althoughin many other respects these molecules are similar(see Figure 4-7 for the PTCBI unit cell structure).Hence, a combined study of these two compounds wasused to develop a model with which to study therelationship of charge delocalization to quantum sizeeffects.135 Finally, the electroabsorption spectra ofPTCDA and PTCBI thin films were measured toobtain a direct measurement of the exciton radiusin order to quantitatively understand effects observedin the thinnest films.135,176,177In a series of experiments described by Haskal and

co-workers,135,163 the low-temperature absorption andfluorescence spectra of alternating multilayer stacksof PTCDA and NTCDA, or PTCBI and NTCDA layerswere measured, where the total thickness of all ofthe PTCDA or PTCBI layers was 500-600 Å. Thethickness of the intermediate NTCDA layers matchedthe PTCDA or PTCBI layer thicknesses, except forthe 50-period 10 ( 5 Å PTCDA structures. In thatcase, the NTCDA thickness was increased to 20 Å toensure growth of a continuous layer. NTCDA wasused as the intermediate layer since its lowest energyexciton state is ∼1 eV higher than the excited stateenergies in PTCDA and PTCBI.203 These multilayerstructures are therefore similar to inorganic semi-conductor multiple quantum well (MQW) structures,where the quantum well material is either PTCDAor PTCBI and the “larger bandgap”, or barriermaterial, is NTCDA.Fluorescence spectra, obtained by optically pump-

ing the multilayer stacks of alternating layers of

PTCDA and NTCDA at a wavelength of 488 nm froman Ar ion laser at T ) 4.2 K, are shown in Figure4-8. Four equally spaced peaks (A, B, C, and D) aredue to vibronic transitions from the excited to theground state. These peaks correspond well to theCTST f S0 transitions: CT[ST-3], CT[ST-2], CT[ST-1], and CT[ST-0], respectively. The energy spacings,ωph, of the vibronic peaks are observed to increasefrom 1065 cm-1 at 500 Å to 1226 cm-1 for thethinnest, 10 Å layers. These vibronic energies aresimilar to those observed in the fluorescence spectraof PTCDA solutions discussed in section 4.1.1, andlisted in Table 7. Taking this, along with the directmeasurement of the Raman spectrum of PTCDA,these dominant features are identified as the C-Cbond stretch (measured41 by Raman scattering at1061 cm-1).The change in ωph (or ∆ωph) in PTCDA films as a

function of layer thickness, as referenced to the 500Å film, is shown by the solid squares in Figure 4-9.For each measurement, ∆ωph is found by fitting theappropriate fluorescence spectrum to a series of fouroverlapping Gaussian curves using parameters simi-lar to those used to fit the solution spectra of PTCDA(see section 4.1.1, Table 7)sa procedure which yieldsboth the peak position and width. The increase inωph with decreasing layer thickness was observed tobe accompanied by an increase in the fluorescenceintensity of the high-energy transitions at the ex-pense of the low-energy transitions. The fluorescenceintensity of the 10 Å film is approximately 12 timeslarger than for the 500 Å PTCDA film, indicating thatinterfaces assist in the radiative recombination of

Figure 4-7. Perspective views of a PTCBI unit cell. Top:Molecular formula for PTCBI (from ref 135).

Figure 4-8. Low temperature (4.2 K) fluorescence spectraof PTCDA/NTCDA multilayer stacks for layer thicknessesof PTCDA ranging from 10 to 500 Å. The four peaks in thespectra are labeled A, B, C, and D, and the spectra aredisplaced vertically for clarity. These peaks correspond tothe following CTST f S0 transitions shown in Figure 4-5:CT[ST-3], CT[ST-2], CT[ST-1], CT[ST-0], respectively (fromref 135).

1846 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 55: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

excitons in PTCDA layers, in the absence of a highdensity of interfacial defects, which would be ex-pected to lead to fluorescence quenching. This in-creased radiative recombination is characteristic ofexciton spatial confinement due to an increasedoverlap of the electron and hole wave functions (seesection 4.2.3.2).The low-temperature (4.2 K) absorption spectra of

a single 500 Å thick PTCDA layer and a 50-periodstack of 10 ( 5 Å PTCDA/20 Å NTCDA are shown inFigure 4-10. A graph of the derivative of the absorp-tion with respect to photon energy, (dR/dE), in theenergy range of the CT exciton peak is shown in theinset.135 Note that the CT peak at 2.234 eV is shiftedto higher energy (as shown clearly in the zero-crossing of each derivative curve) as the layer thick-ness is decreased. A similar blue shift in the CT peakin PTCDA/InPc-Cl multilayer structures has alsobeen observed.204 In that case, an energy increaseof ∼80 meV was obtained for a decrease in PTCDAthickness from ∼250 to 3 Å, as shown in Figure 4-11.These shifts have been attributed to the change inthe CT exciton binding energy as a result of quantum

confinement in a finite potential well.23,24 The po-tential well width corresponds to the PTCDA layerthickness, with the energy barriers provided by theHOMO offsets at the heterointerfaces with NTCDAlayers (or possibly in the case of PTCDA/InPc-Cl, bythe 2 ML thick phthalocyanine layer). Furthermore,no apparent shifts in the higher energy exciton linesare observed, possibly as a result of the Frenkelorigin of these lines (see section 4.1.1).Quantitative fits to the PTCDA/NTCDA data of So

and Forrest,24 and later by Shen and Forrest176 foundthat the radius of the CT exciton in PTCDA is ∼12Å, or 3-4 molecules in the stacks of PTCDA normalto the substrate. This is consistent with spectroscopicassignments of this feature discussed in section 4.1.1,where the CT exciton is found to be an extended stateexisting only in macroscopic aggregates and thinfilms. The dependence of the CT exciton energy shiftwith well width is shown by the data points in Figure4-12. The solid line is a fit to the data (see section4.2.3.2) assuming that the shifts arise from CTexciton confinement in PTCDA layers whose dimen-sions are comparable to the exciton radius of 12 Å.

Figure 4-9. Measured change in the ground state vibra-tional frequency, ∆ωph, of both PTCDA (squares) andPTCBI (open circles) as a function of layer thickness, asreferenced to their respective 500 Å films. The solid line isa fit to data of the observed exciton energy shift (∆ωex) inPTCDA layers using the theory in ref 24 (from ref 135).

Figure 4-10. Low temperature (4.2 K) absorption spectraof a 500 Å PTCDA film and a 10 ( 5 Å PTCDA/20 ÅNTCDA multilayer stack. Inset: Change in absorptioncoefficient as a function of wavelength, dR/dλ, for thesamples in Figure 4-8 (from ref 135).

Figure 4-11. Absorption spectrum for PTCDA/InPc-Clmultilayers grown on sapphire showing the spectral shiftin the low energy PTCDA (straight line). Also shown is thesolution spectrum of PTCDA corresponding well with thatin Figure 4-3 (from ref 204).

Figure 4-12. Energy shift of the PTCDA CT absorptionpeak, ∆E, as a function of PTCDA layer thickness d. Thesymbols correspond to the experimental data, and the solidcurve is the fit using mh,⊥ ) 0.16 mo. The fit is alsocompared with the dashed curve usingmh,⊥ ) 0.18mo andthe dotted line using mh,⊥ ) 0.14 mo (from ref 176).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1847

Page 56: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

The low-temperature fluorescence spectra of mul-tilayer stacks of PTCBI layers alternated with NTC-DA layers, where PTCBI layer thicknesses werevaried from 20 to 500 Å, were used for comparisonto the PTCDA/NTCDA data.135 In the case of PTCBI,only two fluorescence peaks were resolved at 1.478eV and 1.427 eV. In contrast to the substantialspectral changes in the PTCDA spectrum shown inFigure 4-8, the only change observed in the PTCBI/NTCDA fluorescence spectrum as the PTCBI layerthickness was decreased to a 25-period stack of 20 ÅPTCBI alternated with 20 Å NTCDA was a smallredistribution of fluorescence intensity to higherenergy, and a peak width broadening attributed tointerface roughness in semiconductor (e.g., GaAs/AlGaAs) MQWs.205 The change in spacing betweenthe two peaks in the PTCBI fluorescence spectra asa function of layer thickness is shown by the opencircles in Figure 4-9. If it is assumed that the peaksare due to the same intramolecular vibration, thenthese data suggest that the vibrational frequency inPTCBI does not change significantly as layer thick-ness is decreased (∆ωph ≈ 0).The fluorescence efficiency of PTCBI was found to

increase by only a factor of 3 from the widest (500 Å)to the narrowest (20 Å) layers, as compared with the12-fold increase in PTCDA fluorescence efficiency aslayer thickness was reduced over a similar range.This indicates that neither the presence of hetero-interfaces nor quantum size effects in the PTCBImultilayer stacks are as active in enhancing radiativerecombination as they are in the PTCDA multilayerstacks. Furthermore, no absorption peak shifts wereobserved in the PTCBI/NTCDA multilayer struc-tures. This result is consistent with electroabsorp-tion measurements for PTCBI,135 where only small-radius (a* e 3 Å), Frenkel excitons were identified,which are unlikely to exhibit confinement effects forthe range of layer thicknesses used in these experi-ments.To quantitatively understand the spectral depen-

dence on layer thickness, Haskal et al. assumed thatPTCDA could be approximated as an adiabaticsystem with a single, totally symmetric vibrationalmode dominant in the ground state parabolic poten-tial. This assumption provides a single value of ωphbetween the PTCDA vibronic transitions CT[ST-νn]in Figure 4-8. Figure 4-13 shows the modifiedmolecular potential of Figure 4-5 of both a thin(Figure 4-13a) and a thick (Figure 4-13b) PTCDAlayer, indicating the optical transitions between theexcited and ground states in these films. The ob-served redistribution of intensity to higher energyvibronic transitions was attributed135 to a change inthe relative equilibrium configurational coordinateof the excited (qm′) and ground states (qmo), where ∆qm) qm′ - qmo decreases with layer thickness (Figure4-13). In the Condon approximation, the oscillatorstrengths of the transitions from the lowest vibra-tional level of the upper (excited) manifold to the nthexcited vibrational level of the lower (ground) mani-fold can be described by a Poisson distribution.206 Thevibronic transitions in Figure 4-8 have been fit tosuch a distribution135 to obtain ∆qm as a function oflayer thickness, d shown in Figure 4-14a, assuming

a vibrational reduced effective mass of mr ) 6 amuconsistent with a C-C vibrational mode. Here, ∆qmis found to continuously decrease with PTCDA layerthickness over the range of 500-10 Å. The magni-tude of ∆qm corresponds to a large molecular confor-mational change upon excitation which must takeplace in PTCDA, with its exceptionally large Franck-Condon energy, broad absorption spectrum, and lackof mirror symmetry between fluorescence and ab-sorption (see Figure 4-2b).Since ∆qm corresponds to the total molecular con-

formational change upon excitation, these results

Figure 4-13. Configuration coordinate diagram showingthe relative changes in the excited and ground statemolecular potential energy wells of (a) thin and (b) thickPTCDA films (from ref 135).

Figure 4-14. (a) Change in the relative equilibriumconfigurational coordinate (∆qm) of the excited and groundstates as a function of layer thickness for both PTCDA/NTCDA and PTCBI/NTCDAmultilayer films. (b) Franck-Condon energy in PTCDA and PTCBI, measured as thedifference from the CT[O-F] absorption peak to the firstmoment of the fluorescence spectrum. In both a and b, linesare drawn as a guide to the eye (from ref 135).

1848 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 57: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

suggest that the decrease in ∆qm with layer thicknessis due to exciton confinement in ultrathin layers.That is, as the layer thickness is decreased, theexciton overlaps with fewer molecules, so that thetotal conformational change of the excited moleculesis reduced. Similarly, a reduced exciton radius in theultrathin layers results in a decreased dipole mo-ment, which in turn decreases the observed confor-mational change. This is consistent with previousstudies of the absorption spectra of organic/inorganicmultilayer stacks, where it was also suggested thatthe intermolecular interaction decreases with layerthickness.144The molecular potential diagrams shown in Figure

4-13 was also used to understand PTCBI/NTCDAmultilayer stacks. The negligible change135 in ωphsuggests that the curvature of the potential is unaf-fected in MQW structures with even the thinnest (20Å) layers. To account for the small redistribution offluorescence intensity between the peaks in thefluorescence spectra, ∆qm for PTCBI was also calcu-lated. With these assumptions, ∆qm ≈ 0.17 Å, orroughly one-third that of PTCDA (Figure 4-14a), andshows only a minimal dependence on layer thickness.Figure 4-14b shows the Franck-Condon energy

(EFC) in both PTCDA and PTCBI as a function oflayer thickness. Here, EFC is measured from thelowest energy absorption peak to the first moment135of the fluorescence spectrum. It has been shown thatthe exciton-phonon coupling is27,206

which is plotted for both PTCDA and PTCBI inFigure 4-15. Here, p is the Planck’s constant dividedby 2π. The exciton-phonon coupling is found to beindependent of layer thickness. This result is ex-pected for layers whose structure is independent oflayer thickness, which is found to be the case forPTCDA and PTCBI layers grown by OMBD using theconditions described in section 3.3.2. Indeed, Fenteret al.111 have conclusively shown that strain is notrelieved, even for very thick layers (>70 ML) whenmolecular films are grown under nonequilibrium (i.e.,QE) conditions, such as is the case in the experimentsdiscussed here (see Figure 3-43a). Hence, strain

relief is not the source of the observed spectral shifts.Finally, γ in PTCBI is ∼30% smaller than forPTCDA, consistent with the difference observed in∆qm for these materials.From results for PTCDA/NTCDA multilayers, the

effects of confinement are significant due to therelatively large spatial extent of the exciton. Thatis, the extended exciton state (with a radius ∼12 Å,see section 4.2) overlaps a large number of PTCDAmolecules in the crystal. As a result, the excitonbinding energy, radius, and dipole moment are allchanged in thin films whose dimensions approach thescale of the unbound, self-trapped CT state. Thesignificant π-orbital overlap in this closely stackedmolecular system has the additional affect of leadingto strong coupling between molecules which resultsin an unusually large exciton-phonon coupling of γ∼ 2300 cm-1. Such a large coupling implies thatchanges in exciton energy must result in changes inthe intramolecular phonon energy, ωph, as observed.These findings for PTCDA are in contrast to PTCBI/NTCDA, where the PTCBI exciton is largely unaf-fected by the ultrathin layers, due to its small radius,Frenkel-like characteristics. Haskal has explainedthe differences between PTCDA and PTCBI asfollows:135 (i) Since the imidazole group in PTCBIacts as an electron donor, holes are easily trappedon the excited molecules, leading to the formation ofFrenkel excitons. (ii) The random distribution of thetrans and cis isomers present in the PTCBI crystalleads to a disruption of the π-orbital overlap betweenadjacent molecules in a stack, thereby limiting thespatial extent of the exciton. (iii) Calculation of theelectronic overlap of the p orbitals of adjacent carbonatoms in two different molecules indicates that theoverlap is over three times greater in the closelystacked PTCDA films as compared with that due tothe 7% larger intermolecular separation in PTCBI.This implies that intermolecular carrier transfer issignificantly reduced in PTCBI due to its relativelylarge interplanar stacking distance.

4.2. CT States in Nonpolar OMCs: Theory andExperimental ConfirmationHaving eliminated extrinsic sources for the ob-

served excitonic spectral shifts, we now consider theintrinsic physical origin of these phenomena. Acomprehensive theoretical foundation has been de-veloped to treat this apparent “exciton confinement”,and has been used to quantitatively analyze both theelectroabsorption (EA) spectra and CT state absorp-tion blue shifts in PTCDA thin films. We will presenthighlights of this treatment in this section.The delocalized Wannier-Mott exciton and the

localized Frenkel exciton have been theoreticallyinvestigated over the past 50 years, and quantummechanical models have been developed to describethese fundamental excitations. In the first case, theWannier-Mott exciton is based on band structure inuncorrelated systems arising from significant overlapof crystalline periodic potentials. Assuming a single-electron model by applying the Hartree-Fock ap-proximation, the Wannier-Mott exciton can be rep-resented by an orthonormal complete set of spatiallycoherent, hydrogen-like wave functions.207-209 Incontrast, the Frenkel exciton is considered as an

Figure 4-15. Exciton-phonon coupling frequency, γ, asa function of layer thickness for both PTCDA/NTCDA andPTCBI/NTCDA multilayer films. Data are obtained fromFigure 4-14 using method described in text. The lack ofvariation in γ with thickness indicates that the structureof the layers is not a function of thickness, consistent withgrowth experiments discussed in section 3 (from ref 135).

γ ) (EFCωph/p)1/2 (4.1)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1849

Page 58: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

intraatomic or intramolecular excitation character-istic of highly correlated systems, which can diffusefrom site to site. Frenkel states have no internaldegrees of freedom and, hence, no internal quantumnumbers. As a result of polarization of the electronicsubsystems of the surrounding lattice due to chargesand excitons (i.e., many-body interactions), polaron-type quasi-particles are created. This renders thesingle-electron approximation inapplicable, and henceis typically replaced by a Hamiltonian210,211 thatdescribes the Frenkel state by the exciton bandnumber and wave vector. However, the internalquantum number of the Wannier exciton may in factcorrespond to the Frenkel exciton band number,while such Frenkel excitons can be regarded as alimiting case of the Wannier state, where the relativeelectron-hole (e-h) motion is confined to a smallregion of atomic or molecular scale.The charge-transfer (CT) exciton state is usually

considered as an unrelaxed polaron pair,210 with boththe positive and negative polarons of the charge pairlocalized on a set of discrete, identifiable, and nearlyadjacent molecules.212,213 The relaxed polaron stateis called a charge pair (CP) state or self-trapped CTexciton,172,197,209 which is at an energy lower than CTstates due to strong exciton-phonon coupling (seeFigure 4-5).This localized picture is true only if each molecule

is a point source of a δ-function-like potential. Inthat case, charges can only exist in a deep, narrowpotential well at each nuclear or molecular site. Ingeneral, however, the charge is not static. And sincethe molecular ion has finite size, shape, and orienta-tion, the potential cannot be accurately modeled bya δ function. In addition, for planar stacking mol-ecules such as PTCDA and BTQBT193,214 where thereis significant overlap between the π systems ofadjacent molecules in the stack, the electronic statescan no longer be considered as highly localized at asingle lattice site. As a result, the exciton wavefunction must be a somewhat uncorrelated, extendedstate with considerable spatial coherence, and whoseshape and symmetry is dependent on the crystallineand molecular structures.215

As shown in the previous section, extended CTstates and quantum exciton confinement in organicthin films have already been proposed and ob-served.23,24,135,150,197 For multiple quantum wells con-sisting of ultrathin layers of PTCDA sandwichedbetween similar layers of NTCDA, the energies of theexciton absorption peak (Figure 4-12) and the vibra-tional frequencies in the PTCDA fluorescence spec-trum (Figure 4-8) increase with decreasing layerthickness. Also, photoemission and inverse photo-emission studies of PTCDA show that the electron-hole exchange energy, Vexc, which is the energyneeded to remove an electron from one molecule andplace it on another, is215,216∼1 eV, or somewhat largerthan the estimated molecular bandwidth of 0.1-0.4eV.171,197,217 This indicates weak correlation resultingfrom significant charge delocalization in PTCDA. Allof these results suggest that extended CT excitonsare best described using quantum (extended) ratherthan simple electrostatic (localized) models typicallyapplied to organic molecular crystals.195,210

Starting from the real-space representation ofchemical bonding, Mukamel and co-workers28,218-220

proposed a second quantization approach to under-stand quantum size effects in ultrathin organic films.Their method is to map the electronic motions of theconfined states onto a set of coupled harmonic oscil-lators. On the basis of the single electron, reduceddensity matrix, and then solving the time-dependentHartree-Fock equation, the electronic charges andmotion can be related to their optical response.220 Byusing the tight-binding Hamiltonian for π electrons,a basis set of exciton states is chosen to correspondto the discrete e-h separations, where each state isperturbed by Coulomb interactions, electron-electroncoupling, lattice polarization, aggregate response,chemical bond oscillations, and interactions whichinduce the intermolecular charge transfer. The rela-tive magnitudes of these terms determine the char-acteristics28,219,221 of the exciton (i.e., whether itfollows Frenkel, CT, or Wannier behavior). By thismeans, it was found218 that quantum confinement ofthe relative (correlated) e-h motion of the CT excitonplays an important role in determining the magni-tude of the third-order nonlinear susceptibility, ø(3),of conjugated polyenes.At low exciton densities where exciton-exciton

coupling can be neglected, the self-consistent fieldapproximation is valid. In this case, the e-h motioncan be described by a quasi-particle interacting withan average field of the remaining charges. Based onthis quasi-particle picture, the main goal of the workof Shen and Forrest176 was to find a simple approach(i.e., a one-electron Hamiltonian) to describe therelatively delocalized CT exciton in anisotropic, non-polar organic crystals, to study the effect of crystal-line anisotropy on CT exciton energies, and to un-derstand the experimental phenomena such as“quantum confinement” in multilayer structures, theelectroabsorption (EA) spectra,177 as well as otherquantum phenomena frequently observed in nonpo-lar organic molecular crystals. This treatment hasbeen found to be quite general, and was also used todescribe the EA spectra in inorganic semiconductorssuch as GaAs.13

4.2.1. Hamiltonian for CT Excitons

The following general form was chosen for theHamiltonian describing the motion of electron-holepairs in OMCs such as PTCDA:176

where T is the kinetic energy, Ve-h is the electron-hole binding energy (or Coulomb energy), and Vpseudo

is the periodic crystalline pseudopotential. For semi-conductors with a large overlap of atomic wavefunctions between ionic cores in the crystal, theperiodic pseudopotential (which includes on-site cor-relation effects) applied on the valence electrons isweak enough to result in a modified free electron gas.In this limit, the motion of the electron can bedescribed by the one-electron Schrodinger’s equation,leading to a description of the energy levels in terms

Ho ) T + Ve-h + Vpseudo (4.2)

1850 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 59: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

of a family of continuous functions, En(k), each withthe periodicity of the reciprocal lattice. This resultsin an incipient band structure, where n refers to theenergy band, and k is the conserved crystal momen-tum. This small lattice periodic pseudopotential,Vpseudo, leads to relatively high carrier mobility (>0.01cm2/(V s)), and wide energy bands (greater than oron the order of the electron-hole exchange energy,Vexc). Figure 4-16a shows a schematic representationof Ve-h, and the total “delocalized” exciton potentialVdeloc ) Ve-h + Vpseudo, typical of ionic or covalentlybonded semiconductors. Under such conditions, theexciton states are hydrogenic, where the lowestenergy (s-like) electron wave function density is(φ*φ)deloc, with its characteristic monotonic decay andextended spatial coherence.In contrast to inorganic semiconductors, organic

molecular crystals are typically characterized byweak intermolecular vdW interactions. This leadsto relatively small overlap of molecular orbitals and,hence, a strong localization of charge carriers. Undersuch conditions, both the one-electron approximationand the concept of extended energy bands are notappropriate. Additionally, elastic and inelastic in-teractions between the charge carrier and lattice arestrongly coupled, such that the wave momentum kis not conserved (leading to poor spatial coherence),thus further increasing charge localization.222,223 Inthis case, Vpseudo is large compared to Ve-h. Here,Vpseudo corresponds to the on-site correlation energywhich approximately equals the difference betweenthe exchange energy, Vexc, and the crystal polariza-tion energy, PE(r). This is characteristic of manyvdW solids, leading to a low charge carrier mobility(<10-4 cm2/(V s)) and narrow (<kT) energy bands.Figure 4-16 also shows the total potential, Vloc, andwave function density, (φ*φ)loc for such “conven-tional”, highly correlated OMCs where the e-h pairis localized on the molecular sites. Indeed, thecalculated band structure for anthracene indicatesthat electronic polarization significantly reduces theeffective bandwidth. The resulting electron mean

free path is found to be smaller than the intermo-lecular spacing, demonstrating the inconsistency ofthe band model for anthracene-type crystals.222,223

However, the experimental fluorescence and ab-sorption data for ultrathin PTCDA films suggest thatfor some molecular crystals where there is significantoverlap between the π-systems of adjacent molecules,the energy bandwidth is comparable to or largerthan220 Vpseudo while the mean free path is at least asgreat as the intermolecular spacing, leading to sig-nificant charge delocalization. Since the electronicpolarization time constant210 is very short (10-16 to10-15 s), the one-electron model might still apply insuch cases.Given the unusually small intermolecular distance

in PTCDA which leads to a high hole mobility (∼1cm2/V-s) along the stacking direction,172,224 spatiallyextended CT states,23,24,135,177 with large oscillatorstrengths and a combined lowest unoccupied molec-ular orbital (LUMO) and highest occupied molecularorbital (HOMO) energy bandwidth of 0.1-0.4eV,171,197,215,217 is comparable to Vpseudo∼ 0.5 eV. Here,Vpseudo ) |Vexc - PE|, with PE ) 1.5 eV for PTCDA.176Using Vexc ≈ 1 eV measured using photoelectron andinverse photoelectron spectroscopy,215,216 we infer thatfor PTCDA and similarly “large bandwidth” vdWsolids, Vpseudo is small and correlation effects do notdominate the electronic behavior of the material.Using these arguments, the Hamiltonian for delo-

calized CT excitons in the center of mass coordinateshas been written as176

Here, r ) re - rh is the displacement vector of thee-h pair, mi ) me,imh,i/(me,i + mh,i) is the reducedeffective mass along direction i ) ⊥ or |, andme (mh)is the effective electron (hole) mass. For PTCDA, ⊥

Figure 4-16. Schematic diagram showing Ve-h (dashed curve) with the periodic correction term (or pseudopotential, Vpseudo),which is a function of the molecular size, shape and separation. The sum of Ve-h and Vpseudo gives the total potential (solidcurve) for an exciton as a function of e-h separation r. Also shown are plots of the wave function density, φ*φ vs r for (a)delocalized and (b) localized exciton states. The lower part of each figure provides a schematic representation of the localizedand delocalized CT exciton states.

Ho ) T + Ve-h + Vpseudo =

- p2

2 (∇|2

m|

+∂⊥

2

m⊥) - q2

r+ PE(r) (4.3)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1851

Page 60: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

is taken along the substrate normal,73 z. Finally, thedisplacement-dependent polarization energy, PE(r),is a function of the crystal symmetry and molecularstructure.Polarization of the crystal due to the presence of

an excess charge is illustrated schematically inFigure 4-17. These effects, which are typically on theorder of 1-2 eV in OMCs, must be calculated to fullyaccount for the binding energy between the electronand hole forming the exciton. Calculations of polar-ization energies using classical electrostatics werecarried out by Silinsh using the self-consistent po-larization field (SCPF) method225 and by Bounds andSiebrand226 in terms of Fourier-transformed latticemultipole sums. They found that, to first order, thee-h interaction potential (Ve-h) is essentially Cou-lombic, e.g., Ve-h ) -q2/εavgr, for large charge pairseparations, r (i.e., r > one lattice constant), whereεavg is the spatially averaged crystal dielectric con-stant. This result is, in some respects, similar to thecase of conventional semiconductors where the crys-tal lattice consisting of electrons and nuclei can beregarded227 quasi-macroscopically as polarizable me-dia with a spatially averaged dielectric constant, εavg,as long as the e-h distance is large compared to thelattice constant. Fundamentally, an electron in theconduction band, or the lowest unoccupied molecularorbital (LUMO), as well as a hole in the valence band,or the highest occupied molecular orbital (HOMO),are always accompanied228 by electronic polarization(via polarons) of the surrounding medium. Thissimilarity suggests the feasibility of applying theWannier exciton model to describe CT excitons withradii larger than one lattice constant.Hence, the crystal polarization in the anisotropic

OMC, PTCDA, as well as in NTCDA was calcu-lated176 using a method similar to that of Bounds etal.226,229 From this analysis, PE(r) in the effective one-exciton Hamiltonian in eq 4.3 can be used to findenergies and wave functions of CT excitons by solvingSchrodinger’s equation. To calculate PE(r), Shen andForrest176 considered a lattice with a basis whose unitcell volume is v, and a molecule at r(lk), where l )lx,ly,lz is the lattice index, and k ) 1,2,3,... is theindex of a particular basis molecule within the unit

cell. The medium is polarized due to the electric field,Fo, from excess charges in the crystal. The macro-scopic field Fmac at site (lk) thus includes Fo and theinduced dipole field, Find via

where Find is the sum of the field due to the dipole atthe same site (self-field, Fself) and that due to thesurrounding dipoles. Assuming a spherical dipole forhomogenous and isotropic media, Fself is taken as230-µ(lk)/3νεo, where ε0 is the vacuum permittivity andµ(lk) is the dipole moment tensor. Then

where R(k) is the effective polarizability of moleculek, and Floc(lk) is the local (average) field at site (lk)excluding the induced field due to molecule (lk).Hence

The total crystal polarization energy, PE, is foundby integrating over the individual molecular polar-izations, pE(y) over the first Brillouin zone (1BZ)

Thus, the e-h interaction energy is a sum of theunscreened Coulomb attraction and the crystal po-larization energy, via

Equation 4.8 gives the potential energy of an e-hpair in a nonpolar organic crystal and includes thee-h attraction as well as the microscopic latticeresponse to the charge pair in terms of the polariza-tion energy, as required in the Hamiltonian in eq 4.3.Equation 4.8 has been used to calculate Ve-h in

NTCDA and PTCDA, using known dielectric tensorelements.175 The crystalline and dielectric anisotro-pies of NTCDA are much smaller than for PTCDAdue to the herringbone stacking habit in NTCDA.This leads to a nearly isotropic index of refractionfor NTCDA, resulting in a spatially averaged diago-nal dielectric tensor whose elements along the prin-ciple axes are εavg ) 2.6.The interaction energies, Ve-h(r) for both NTCDA

and PTCDA are shown in Figure 4-18. The resultsare compared with the “screened” Coulomb potential,Vscr ) -q2/|ε‚r|, where ε is the macroscopic dielectrictensor for the crystal. Since diagonal elements of theNTCDA dielectric tensor are approximately equal,176the e-h interaction energy in Figure 4-18a is com-pared with Vscr ) -q2/εavgr, providing a good fit tothe calculated values along all crystalline axes.For the highly anisotropic PTCDA, Ve-h depends

sensitively on the crystal direction for r < 20 Å. Thedashed curve in Figure 4-18b is a plot of Vscr ) -q2/εavgr, where εavg ) 3/(1/ε⊥ + 2/ε|) ) 3.2. Using εavg

Figure 4-17. Schematic representation of lattice polariza-tion in an OMC. The cross-hatched circles represent themolecular positions within the lattice, and an excess chargeis placed at (-) in the center of the diagram.

Fmac(lk) ) Fo(lk) + Find(lk) (4.4)

µ(lk) ) R(k)‚Floc(lk) (4.5)

Floc(lk) ) Fmac(lk) - Fself(lk) (4.6)

PE ) -1/2∑kFo(lk)‚µ(lk) ) -v/2∫1BZpE(y) dy (4.7)

Ve-h ) - q2

|r(l1k1,l2k2)|+ PE[r(l1k1,l2k2)] (4.8)

1852 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 61: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

results in the best fit to Ve-h for large r (>25 Å),regardless of direction, suggesting that the potentialbecomes spatially averaged at distances of only a fewmolecular lattice constants from the origin, even foranisotropic systems such as PTCDA. However, forsmall r (<15 Å), Ve-h deviates significantly from thespatial average (clearly apparent in the inset ofFigure 4-18b, where Ve-h is plotted as a function of1/r). For small r, the mean dielectric constant canbe approximated using Vscr ) -q2/εir (where i refersto crystalline directions), as shown by the corre-sponding curves in Figure 4-18b.Once Ve-h was determined, a trial wave function

was chosen to solve eq 4.3. Due to the relativeisotropy of PTCDA in the | direction, cylindricalcoordinates (r ) (F,φ,z) with F, φ in |, and z in ⊥directions) were used. In these reduced coordinates,the kinetic energy operator is isotropic, where theanisotropic mass and dielectric tensor elements areexpressed in the parameter, âi ) (mi/mo)1/2 εavg/εi. Heremo is the free electron mass and mi is the reducedmass in the i direction. The trial wave function

chosen for the lowest, 1s exciton state was231

where

Here, c| and c⊥ are variational parameters used tominimize energy, F′ ) εiaBF, and z′ ) εaBz, and aB )0.529 Å is the Bohr radius of the free electron. Thelowest energy, E1s

(0), is found via

and the exciton radius along i is then

4.2.2. Theoretical Fits to Electroabsorption and MQWAbsorption Spectra4.2.2.1. Electroabsorption Spectrum. The theory

outlined above was used to fit the EA spectrum ofPTCDA. Here, EA has been shown to be an ex-tremely useful means for providing information aboutband structure and localized states in both organicand inorganic semiconductors.13,177,212,232,233 Localizedstates in OMCs result in correspondingly smallerelectric field effects than in inorganic semiconductorswhere the low exciton binding energy (∼10 meV) canbe strongly perturbed by even small external fields.In OMCs, the first- and second-order Stark effectsare typically interpreted as due to a localized dipolemoment interacting with the applied electric field.212,234For tightly bonded (∼1 eV) Frenkel excitons, thechange of absorption follows the first derivative ofthe absorption with respect to energy (i.e., thequadratic Stark effect), whereas for polar CT excitonswith large transition dipoles and with binding ener-gies e500 meV, the field-induced change of absorp-tion is proportional to the second derivative of theabsorption spectrum.212,234 For example, electrore-flectance studies of the π-π* excitations of thepolydiacetylene single crystal poly-1,6-di-N-carba-zolyl-2,4-hexadiynal (DCHD) provided evidence foran extended CT exciton state whose EA response is10 times larger than for the localized Frenkel excitonlines in that same material.212,233,235The room temperature excitonic EA spectrum due

to excitons in crystalline thin films of PTCDA hasbeen studied177 at electric fields ranging from F )|F| ) 100 to 300 kV/cm . This field is sufficientlysmall to consider its effects to be a linear perturbationon the exciton binding energy. Given Ho, along withthe basis wave functions, æn, provided in eq 4.13, theEA spectrum of PTCDA has therefore been fit usinga perturbation Hamiltonian, H1. Since an externalfield Fo will polarize the crystal, the local field Flocat the exciton, and the macroscopic field Fmac in thecrystal, may differ. Hence

Figure 4-18. Electron-hole potential energy (Ve-h) calcu-lated for (a) NTCDA and (b) PTCDA as a function of chargeseparation (r) along three crystalline axes (data points).Since NTCDA is isotropic, the calculation is compared withresults obtained in the isotropic continuum approximation,where ε) 2.6 (solid line). The calculation for the anisotropicPTCDA is carried out along three crystalline axes (a axis,circles; b axis, squares; c axis, triangles). The dashed-dotted curve is Ve-h ≈ Vscr using the dielectric constantalong the a axis, εa ) 21, and the dotted curve uses valuealong the b, c, εb ) 4.5 (where εb ≈ εc). The dashed curveuses the spatial averaged dielectric constant εavg ) 3.2. Notewhen r > 25 Å, εavg results in the best fit, and when r < 15Å, εi, where i represents the crystalline axes, providesexcellent approximations to the calculated values. The solidcurve uses the effective dielectric constant, εeff describedin text (from ref 176).

æ1s ) exp(-g)/xπc|2c⊥ (4.9a)

g ) (F′2/c|2 + z′2/c⊥

2)1/2 (4.9b)

Eh 1s(0)(cj|,cj⊥) ) min

c|,c⊥⟨æ1s|Ho|æ1s⟩

aji ) aBcjiεavgxmo/mi

H1 ) q∫rerhFloc dr ) qFmac‚r (4.10)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1853

Page 62: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

This treatment is valid only at low electric fields,where qFaavg ) q|F|aavg < E(0)

1s = 200 meV, or F < 2MV/cm, where

Using the perturbed trial wave function236

the energy shift in the CT absorption peak (∆E) wasfound for a field oriented in the normal (⊥) and in-plane (|) film directions, as

Also, the broadening of the state due to the externalfield is given by

The ratio of the exciton line shift broadening istherefore independent of crystal direction:

In the limit of small fields, the change of the excitonabsorption spectrum depends quadratically on theelectric field, via

where Eph is the photon energy. The experimentalquadratic dependence of ∆R is shown in the inset ofFigure 4-19 for PTCDA,177 where the electric field

was applied along ⊥, and ∆R was measured at theCTST and CTF energies of Eph ) 2.1 and 2.2 eV,respectively (Table 7). The EA spectra obtained forthe electric field oriented along ⊥ and | directions areplotted in Figure 4-20.The lowest energy absorption peak at 2.23 ( 0.03

eV in the PTCDA spectrum was shown in section4.1.1 to be a spatially extended CT state, whereasthe higher energy absorption peaks correspond toFrenkel states.197 Due to the localized nature of theFrenkel excitons, these later states are less sensitiveto the applied electric field.237 Furthermore, sinceFrenkel excitons respond to the electric field via thechange in polarizability, the EA spectra of such statesare proportional to the first derivative of absorp-tion212, ∂R/∂E. Hence, the high energy features (>2.3eV in PTCDA) exhibit a smaller electric field depen-dence and different line shape as compared to thelow energy CT states. The EA spectrum measuredfor PTCDA, shown in Figure 4-20, is thereforeprimarily due to line broadening of the lowest energy,extended CT state.Hence, ∆R was fit by Shen et al.177 using two

parameters ∆E and ⟨H12⟩, while ∂R/∂E and ∂2R/∂E2

were obtained from the lowest, inhomogeneously(phonon) broadened CT absorption line with a peakenergy of 2.23 ( 0.03 eV and a width of 0.17 ( 0.02eV (see Table 7). According to eq 4.14, broadeningdominates the EA spectra due to the small excitonradius (∆E/⟨H1

2⟩ ) -(0.14 ( 0.02)× 105 V/cm2/meV2).Thus, the ratio ∆E/⟨H1

2⟩ is used to estimate aavg, while⟨H1

2⟩⊥/⟨H12⟩| gives the ratio of the exciton radii along

⊥ and | directions. Figure 4-20 shows the EA dataand fits (dashed lines) employing the theoreticaltreatment of section 4.2.1 for normal and in-planefields,176 obtained using aj⊥ ) 12.5 ( 0.5 Å and aj| )10.2 ( 0.4 Å, indicating that the CT state does indeedextend over several molecular sites, consistent withthe analysis of the solution fluorescence discussed in

Figure 4-19. Room temperature PTCDA electroabsorptionspectra at various electric fields oriented normal to thesubstrate, F⊥. Inset: Electroabsorption magnitude as afunction of F⊥

2 at two different photon energies, Eph ) 2.1and 2.2 eV (from ref 176).

Figure 4-20. Absorption (R) and electroabsorption (∆R)spectra of PTCDA. Top: Electroabsorption spectra andnumerical fits (dashed lines) with fitting parameters listedin Table 8. For the field oriented parallel to the substrate,the electroabsorption (circles), ∆R|, was obtained at F| )0.16 MV/cm, and for perpendicular fields (squares), ∆R⊥,the field was F⊥ ) 0.15 MV/cm. Bottom: Absorptionspectrum (triangles) and the fit (lines). The fit of the totalspectrum assumes a series of Gaussian line shapes corre-sponding to previously resolved (ref 197) exciton line shapesfor PTCDA. The dashed-dotted line corresponds to the CTabsorption component of the fit with a peak energy of 223( 0.03 eV and a width of 0.17 eV (from ref 176).

aavg ) x3aj⊥aj|2 ≈ 11 Å

ψ1s ) æ1s1 -2H1εavgaavg

q2 [1 +g1s2 ] (4.11)

∆E ) E1s - E1s(0) ) - 9

4εavgF⊥

2aj⊥2aavg

- 94εavgF|

2aj|2aavg

(4.12)

⟨H12⟩ ) ⟨(qFmac‚r)

2⟩ ) q2F⊥2aj⊥

2

q2F|2aj|

2 (4.13)

∆E⟨H1

2⟩) - 9

4εavgaavgq2

(4.14)

∆R(F,Eph) )∂R(Eph)∂Eph

∆E +∂2R(Eph)

∂Eph2

⟨H12⟩ + ...

(4.15)

1854 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 63: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

section 4.1.1. The small anisotropy found for thePTCDA exciton radii suggests a similarly smallanisotropy in the effective reduced mass, which arecalculated to bem⊥ ) (0.17 ( 0.01)mo andm| ) (0.19( 0.01)mo. The exciton binding energy was E1s ) 185( 20 meV. Results from the EA analysis for PTCDAare provided in Table 8.While the apparent isotropies in the exciton mass

and radii appear unexpected for such an anisotropicmedium as PTCDA, EA provides the “Coulombic”effective mass of the nearly symmetric 1s state. Onthe other hand, hole mobility measurements indicatea large asymmetry (with37 µp,⊥ = 106µp,|). Mobility,however, reflects the “transport” effective mass in-ferred from µ ) qτ/m. Since µ ∝ τ, where τ is thecarrier scattering time, which can carry the anisot-ropy in µ due to enhanced scattering along a par-ticular crystalline direction. For the case of PTCDA,the motion of a charge along | is easily scattered bythe large intermolecular potential barriers (for ex-ample, Vpseudo, in Figure 4-16), leading to stronglocalization in that direction. Hence, we expect thatτ| << τ⊥, leading to the very anisotropic mobilityobserved, while the effective mass remains largelyisotropic, since it is primarily due to the electronicdistribution within a given molecule.The analysis of the EA spectra has also been

applied to study176 Wannier excitons in GaAs. Thisis readily apparent by similarities between the EAspectra of PTCDA and GaAs13 shown in Figure 4-21.The broadening observed in GaAs has been qualita-tively attributed13 to field ionization shortening of theexciton lifetime, optical-phonon collisions, etc. How-ever, a simple, quantitative fit can be achieved byapplication of the wave function and potential em-ployed to describe field-induced line broadening inCT excitons in organic films. The EA data in Figure4-21 were fit (dashed line) assuming a ) 70 Å, and

E1s ) 8 meV, consistent with expected values for bulkGaAs.238 This fit to the bulk GaAs EA line shapeusing eq 4.3 suggests that both Wannier and CTexcitons have a common physical origin and can bothbe equally well approximated by hydrogenic wavefunctions broadened in an external electric field.4.2.2.2. Quantum Confinement in Organic Multi-

layers. Evidence for quantum confinement of ex-tended CT states in organic multilayer structures haspreviously been provided by So,23,24 Haskal,135,163 andHong.150 As discussed above, it was observed thatthe lowest energy absorption peak, the ground-statevibrational frequency of the exciton fluorescencespectra, and the exciton photoluminescence peakshift to higher energies as the layer thickness isdecreased. Shen and Forrest176 have applied thegeneral Hamiltonian of eq 4.3 to quantitatively studythese blue shifts in the CT absorption peak inPTCDA, and then compared the resulting excitonradius, effective mass and binding energy with thoseobtained from direct measurements obtained fromthe EA spectra presented in section 4.2.2.1.The source of the confinement in the organic layers

of thickness, d, are the energy discontinuities of theHOMO and LUMO bands at the heterointerfacesbetween the adjoining materials. The Hamiltonianof an e-h pair in the PTCDA/NTCDA multiplequantum well structure shown in Figure 4-22 hasbeen expressed as176

where T| is the in-plane kinetic energy of the excitoncenter of mass, me-h,| is the reduced mass along |,and Te,⊥ and Th,⊥ are the kinetic energies along thenormal (i.e., molecular stacking) direction with cor-responding massesme,⊥ andmh,⊥. Also, ze and zh arethe normal coordinates for the electron and hole,respectively, and Ve, Vh are the LUMO and HOMOlevels in Figure 4-22. For multilayers consisting ofPTCDA and NTCDA, energy barriers to carrier andenergy transport are formed by the offsets of theLUMO and HOMO energy band minima. Theseenergies were found23 via analysis of the current-voltage characteristics of PTCDA/NTCDA hetero-junctions, to be ∆ELUMO ) 980 meV and ∆EHOMO )50 meV (where NTCDA has the larger HOMO-LUMO energy gap). Since ∆ELUMO . ∆EHOMO, andme . mh, the quantum well for electrons was ap-proximated as having an infinite depth, with theelectron confined in a single layer of “index” N ) 0.In this analysis, only the lowest, nearly symmetric1s CT state corresponding to the lowest energyabsorption peak observed in the PTCDA spectrumwas considered.The trial wave function of the lowest exciton state

for an electron in layer 0, and a hole in layer N can

Table 8. Summary of the Result for PTCDA (From Ref 176)

symbol unit EA results QC results

exciton binding energy E1s meV 185 ( 20 177 ( 20exciton radius along the normal direction a⊥ Å 12.5 ( 0.5 12.9 ( 1.0exciton radius along the in-plane direction a| Å 10.2 ( 0.4effective reduced mass along the normal direction m⊥ mo 0.17 ( 0.02 0.16 ( 0.02effective reduced mass along the in-plane direction m| mo 0.19 ( 0.02

Figure 4-21. Top: Fit (dashed line) to the electroabsorp-tion spectrum (squares) of GaAs at F ) 8 kV/cm taken fromref 13 using the theory in text. Bottom: Band edgeabsorption spectrum of GaAs indicating the Wannierexciton state. Data also from ref 13 (from ref 176).

H ) T| + Te,⊥ + Th,⊥ + Ve-h + Ve(ze) + Vh(zh)

(4.16)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1855

Page 64: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

be written as the product of the free exciton wavefunction (φe-h(F, z ) ze - zh)), and the electron (φe(ze))and hole (φh(zh)) envelope functions. These latterfunctions are oscillatory in the wells and decayingin the barriers. The total wave function is

where

and

withN ) even (odd) corresponding to the layer indexfor wells (barriers), and A(N) is a constant dependingon both g and the effective masses in the ⊥ direction.Thus, k and κ are obtained from solutions of the time

independent Schrodinger equation. Minimizing E1sthe CT[O-F] absorption peak shift data obtained bySo24 and Haskal135 in multilayer structures, whered was varied from 10 to 500 Å, have been fit176 toobtain the exciton radius and reduced masses in the⊥ direction.The results obtained from this fit for different hole

masses are shown in Figure 4-12, where the dashedcurve corresponds to mh,⊥ ) 0.18mo and the dottedline corresponds tomh,⊥ ) 0.14mo. The best fit (solidline) is obtained formh,⊥ ) (0.16 ( 0.01)mo. The blueshift of the exciton energy vs d tapers off for d < 10Å due to the onset of tunneling between wells, inwhich case the holes are no longer further confinedas d f 0. This condition corresponds to that observedfor organic bilayers reported by Hong et al.150The effective mass, radius (a⊥ ) 12.9 ( 1.0 Å) along

the normal direction, and exciton binding energymeasured using the absorption blue shift data forPTCDA/NTCDA MQWs are consistent with the re-sults obtained independently from fitting the EA datafor homogeneous PTCDA thin films (see Table 8).These data provide the first conclusive evidence thatWannier-like CT excitons exist in molecular com-pounds characterized by closely spaced planar mol-ecules, resulting in a large molecular wave functionoverlap between adjacent molecules with extended,Bloch-like, spatially coherent electronic states.This situation is analogous to the directly observed

transition from Frenkel to Wannier excitons in pres-surized fluid xenon.239 Furthermore, the narrowingof the optical bandgap of PTCDA under hydrostaticpressure,240 shown in Figure 4-23, supports thedelocalized picture for the CT exciton in this material.That is, the π-π* transition energy is found tosignificantly decrease with pressure due to thebroadening of the energy bands on close approach ofthe adjacent molecules in the stacks.To further test the effects of quantum confinement,

room temperature, time-resolved fluorescence mea-surements of the CT[ST-2] transition peak at E )

Figure 4-22. Top: Schematic diagram of a PTCDA/NTCDA MQW structure, where N is the layer index. TheHOMO-LUMO energy diagram of this structure is shownto the right, and in greater detail at the Bottom. Also shownin the lower band diagram are schematic illustrations ofconfined wave functions (solid lines) and in the bulk(dashed lines).

ψ ) φe(ze)φh(zh)φe-h(F, z ) ze - zh) (4.17a)

æe(zz) ) cos(π0ze/d) |ze| e d/20 |ze| > d/2 (4.17b)

æh(zh) ) cos[k(zh - Nd)] N ∈ evencos(kd/2)cosh[κ(zh - Nd)]

cosh(kd/2)N ∈ odd

(4.17c)

æe-h(F,z ) zh - ze) )

A(N)exp(-g) N ∈ evenA(N)exp(- mh,b,⊥g

mh,w,⊥) N ∈ odd (4.17d)

Figure 4-23. The low-energy absorption tail shift as afunction of pressure for NTCDA and PTCDA films. Upperstraight line is the same plot for NTCDA vs relative density(from ref 240).

1856 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 65: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

1.70 eV were carried out by optically pumping theMQWs with a 4 ns pulsed Ar ion laser.24 At thisenergy, excitons were generated only in the PTCDAlayers since NTCDA is transparent at this wave-length. The fluorescence decay transients for twoMQW samples obtained at 20 K are shown in theinset of Figure 4-24. Here, the exponential decayover several decades in intensity of the transient isapparent. In addition, the dependence of the excitonlifetime (τex) on well width (d) is shown by data pointsin Figure 4-24, with τex decreasing from 10.8 ( 0.5ns for d ) 200 Å, to 5.7 ( 0.5 ns for d ) 10 Å. Now,the total recombination rate is τ-1 ) τex-1 + τnr-1,where τnr is the time constant for nonradiative decay,such as due to trapping at the PTCDA/NTCDAinterfaces. If the nonradiative recombination rate issuch that τex-1 < τnr-1, the lifetime should decreasedue to traps, thereby decreasing the intensity of theradiative emission. The time-integrated room tem-perature fluorescence intensity, however, was foundto be independent of layer thickness,24 indicating thatnonradiative recombination is not significant. On thecontrary, Haskal135 found that the integrated fluo-rescence intensity significantly increased with de-creasing d (Figure 4-8). A similar inference can bemade from experiments34 which indicate that carriercollection after photogeneration is enhanced in PTC-DA/vanadyl-oxide Pc (VOPc) multilayer structures.In that case, nonradiative centers at the multipleinterfaces would tend to decrease the carrier collec-tion efficiency, in stark contrast to observation.Finally, if the exciton lifetime is dominated by traps,it should decrease with increasing temperature.241However, time-resolved measurements have shownthat the radiative decay time of OMBD-grown PTCDA/NTCDA MQWs is independent of temperature from20 to 295 K.The decrease of τex with decreasing d, therefore,

results from an increase in the overlap of the electronand hole wave functions in the PTCDA potentialwells. A similar explanation of the dependence of theexciton lifetime on well width has been used toanalyze radiative decay in GaAs/GaAlAs quantum

wells.242 For the hydrogenic 1s state, τex-1 is propor-tional to the exciton volume.243 Using the quantummechanical treatment of the CT state, the excitonvolume (∼∫|ψ|2 d3r) in a MQW decreases with wellwidth, thus accounting for the decrease in excitonlifetime. On the basis of the above analysis, thedependence of exciton volume, and hence τex on d, isshown by the solid line in Figure 4-24 using theexciton radius and effective mass provided in Table8. These results are in good agreement with thelifetime data, indicating the exciton quantum-con-finement model can quantitatively account for theobserved changes in both the exciton binding energyand lifetime.

4.2.3. Correspondence between MQW Absorption andFluorescence SpectraThe analyses of the EA and the fluorescence

spectral dependencies on layer thickness have shownthat spatially extended CT excitons exist in PTCDA,while Frenkel excitons dominate the spectral featuresin the similar compound, PTCBI. In addition, analy-sis of the fluorescence spectra has shown that theamount of molecular conformational change uponexcitation decreases with layer thickness in ultrathinlayers of PTCDA (Figure 4-14a). The large changein dipole moment which accompanies the creation ofan excited state244 indicates strong CT exciton-phonon coupling. To quantitatively investigate theexciton-phonon interactions in strongly coupledOMCs such as PTCDA, and in particular to under-stand why the absorption blue shift (∆ωex) andchange in phonon frequency (∆ωph) with layer thick-ness are similar in magnitude (Figure 4-9), Haskalet al.135 analyzed the interaction of excitons andphonons in external radiation fields. The Hamilto-nian used to calculate linear exciton-phonon inter-actions was27

where pωex is the exciton binding energy, ω is thephoton frequency, pωph,â is the energy of the phononmode â, and µ is the electric dipole moment of theexciton. Further, a+(a) and bâ

+(bâ) are the excitonand phonon creation (annihilation) operators, respec-tively.The exciton population (a+a) and phonon amplitude

(bâ+ + bâ) are calculated in steady state assuming

the external field F is small, and only a singledominant phonon mode is present, in which case

and

In eq 4.19a, Ω ) µF/2p is the Rabi frequency, δ isthe exciton dephasing rate, and η ) (ωex - ω) is the

Figure 4-24. Exciton lifetime as a function of PTCDAlayer width in symmetric PTCDA/NTCDA multilayerstructures. The solid line is a fit to the data based on thetheory of So (ref 24) and Shen (ref 176). Inset: Photolu-minescence time decay transients for two MQW samples.The transient for the d ) 200 Å sample has been shiftedupward by 0.1 decade in this plot for clarity (from ref 24).

H ) p(ωex - ω)a+a + ∑â

pωph,âbâ+bâ -

∑â

pγa+a(bâ + bâ+) - µa+F - µaF* (4.18)

a+a = Ω2

η2 + δ2(4.19a)

b + b+ ) 2γa+aωph

(4.19b)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1857

Page 66: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

detuning parameter. In general, δ , ωex. From eq4.19, the exciton-phonon interaction energy is135

Assuming that Eint is independent of changes in EFCor Eex which accompany a reduction in layer thick-ness, and using ∆η ) ∆ωex, then

Thus, the change in the exciton binding energy (∆Eex) p∆ωex) is directly proportional to the change in thephonon frequency (∆ωph) with layer thickness. Here,Eint corresponds to the polaron energy195,245 whichfrom the fluorescence spectra of PTCDA MQWs, isfound to decrease with layer thickness. This indi-cates that the exciton in the ultrathin layer inducesa smaller lattice distortion in the thinnest films.However, Haskal135 found that ∆ωph (and thus ∆EFC)for PTCDA is ∼25% of the value predicted by eq 4.21(the measurement is shown by the solid line, Figure4-9). Since Eint ≈ µ4 (c.f., eq 4.20), this unexpectedlysmall value of Eint suggests that the dipole momentdecreases by ∼10% as the layer thickness is de-creased from 500 to 10 Å. Since µ is proportional tothe exciton radius, this implies that a is also de-creased by a similar amount as a f 0. This reductionin radius is a consequence of confinement of extendedexciton states in the thinnest layers.231

4.3. Optical and Electronic Properties of OtherOrganic Multilayer SystemsSince the first demonstration of the growth of an

organic MQW structure using ultrahigh vacuumdeposition,23 there have been numerous studies onthe growth24,50,135,138,139,141-143,146,148-151,163,176,183,184,204,246

and theory25-29,176,218,220,221,247-251 of multilayer struc-tures aimed at understanding and exploiting theirunique optoelectronic properties. A considerableinterest in hybrid organic/inorganic heterostructuresand MQWs has also developed for similar rea-sons.144,145,159,160,189,190 In this section, we summarizeresults of studies which have been directed at un-derstanding the physics of fundamental excitationsin organic nanostructures other than the very wellcharacterized perylene and naphthalene-based sys-tems discussed in sections 4.1 and 4.2.Quantum confinement has also been used150,246 to

explain blue shifts with decreasing layer thicknessin the optical absorption and fluorescence spectra ofself-assembled ultrathin alternating bilayers of theconjugated copolymer, poly(1,4-phenylenevinylene-co-A,B-naphthylenevinylene), or co(A,B-NV) and thenonconjugated insulating polymer, poly(styrene-4-sulfonate), or SPS. In that work, both A,B ) 1,4 andA,B ) 2,6 copolymers with differing concentrationsof naphthalene were employed. The copolymer unitsfacially stack when deposited by self-assembly fromsolution onto ultra-smooth, float glass substrates ina layer-by-layer fashion. The facial stacking resultsin significant π system overlap, and hence carrierdelocalization, in the substrate-normal direction (i.e.,perpendicular to the polymer chain axis). By varying

the naphthalene concentration in the copolymer, andthe number of bilayers in the assembly, an ultrathinlayer bounded by air and glass on the opposingsurfaces forms the resulting potential well of “infi-nite” depth to the trapped excitons. Typically, thethickness of each co(A,B-NV)/SPS bilayer was be-tween 12 and 16 Å, again dependent of the naphtha-lene concentration in the co(A,B-NV). Since the wellconsists of a collection of bilayers, it was assumedthat the insulating SPS was sufficiently thin suchthat it did not present a significant tunneling barrierto the charge carriers (holes) with allowed states inthe copolymer layers.Systematic blue shifts with decreasing layer thick-

ness in both the photoluminescence and absorptionspectra of co(A,B-NV)/SPS bilayer assemblies wereobserved. The photoluminescence peak shift is clearlyapparent in the series of spectra shown in Figure 4-25for samples of different compositions and number ofbilayers (n). In Figure 4-26, the peak shift (∆E) isplotted as a function of 1/d2, where d is the totalassembly thickness between air and the glass sub-strate. The binding energy of excitons confined inan energy well of infinite depth is increased by: ∆E) h2/8d2m*, where h is Planck’s constant, and thereduced effective mass, m* is 1/m* ) 1/me + 1/mh.In PPV-like polymers such as co(A,B-NV), me . mh.Hence, straight line fits to the data in Figure 4-26yields the hole effective mass, which is found to rangefrom 0.06mo to 0.09mo for naphthalene in co(A,B-NV),with the naphthalene concentration ranging from 1%to 20%, respectively. Given that the fits to the datausing this simple confinement model are quite goodfor bilayer assemblies of thickness d > 40 Å usingvalues of effective hole mass which are consistentwith those found for PPV-like materials, it was

Figure 4-25. Photoluminescence spectra of self-assembledmultilayers as a function of the number of bilayers, n. (a)20% co-(2,6-NV)/SPS, (b) 20% co-(1,4-NV)/SPS, and (c) 20%co-(2,6-NV)/(SPS+0.4M CaCl2). For comparison, the spectraof 1000 Å thick spincast films is also shown in a and b.Note the blue shift as n is decreased for these latter films.Also, note that by addition of CaCl2 (curve c), the PLspectra are independent of assembly thickness (from ref150).

Eint ) -2EFC[Ω/η]4 (4.20)

∆EFC

EFC) 4

∆Eex

Eex) -

∆ωph

ωph(4.21)

1858 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 67: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

concluded150,246,252 that the observed spectral shiftsresult from changes in the exciton binding energy dueto the presence of energy barriers surrounding thepolymer layers, similar to the case of PTCDA/NTCDAmultilayer assemblies.As a test of this hypothesis, the thickness of the

insulating spacer layers between the conductinglayers was varied. When the spacer layer thicknessexceeded ∼40 Å, the spectral shifts in the absorptionspectrum were not observed. This suggests that withsufficiently thick layers, there is poor delocalizationof the excitons along the stacking direction, therebyeliminating the expected thickness dependence in theabsorption and emission spectra. That is, when thespacer layers are thin (∼20 Å or less), confinementis determined by the entire stack, and not solely bythe thickness or even the flatness of a particular layerin the stack.252 Furthermore, by comparing spectrafrom both bilayers and multiple bilayer copolymerassemblies, molecular conformational effects wereeliminated, suggesting that the blue shifts observedare indeed the result of exciton delocalization alongthe bilayer assembly stacking direction (i.e., normalto the substrate plane).More recently, confinement along the chain axis

has also been claimed253 for the conjugated blockcopolymer, TBA-4 (consisting of poly(2,5 benzox-azole)-block-poly(p-phenylene benzobisthiazole)-block-poly(2,5 benzoxazole)). In this case, confinement isinduced by the HOMO-LUMO gap differences (es-timated as 0.7eV) between the blocks along the chain.While this represents a different geometry to thatoffered by planar stacking molecules, in both casescharge is localized along one spatial dimension, andhence the confinement mechanisms (due to potentialbarriers interposed between wells) are fundamentallythe same.The absorption spectra attributed to excitons in the

organic thin-film component of organic/inorganicmultilayer structures consisting PTCDI and MgF2have also been observed to shift to higher energieswith decreasing organic thin-film thickness.145 TheseOMBD grown structures were found to have consid-

erable interface flatness (as determined by grazingincidence X-ray diffraction, see section 3.4.2), forPTCDI layer thicknesses ranging from 5 Å (orslightly more than 1ML) to 50 Å. Blue shifts in thelowest energy peak in the PTCDI optical absorptionspectrum are clearly observed, as indicated in Figure4-27. Following earlier analyses of PTCDA/NTCDAquantum wells,23 Tokito et al.145 also attribute thisphenomenon to quantum size effects in the PTCDIexciton. However, these authors note that the spec-trum of PTCDI-doped MgF2 films also exhibitschanges when the PTCDI concentration is very high(>20%). This observation does not contradict theconclusion that quantum size effects are responsiblefor the spectral shifts, since aggregate formation atsuch high PTCDI concentrations is anticipated. Suchaggregates would form a distribution whose meanradius would depend on concentration. Hence, sizeeffects are also expected in such composite systems,although the spectral changes for the mixtures mightdiffer somewhat from the more structurally homog-enous, alternating layer structures, as has beenobserved.145In a second series of experiments, Tokito and co-

workers145 studied CuPc/MgF2 multilayer structures(section 3.4.2). While a redistribution of energybetween spectral peaks in the CuPc Q bands wasfound on varying the layer thickness from 13 to 50Å, there were no clear spectral shifts that could beattributed to quantum size effects. This finding issimilar to that obtained for PTCBI/NTCDA multi-layer stacks135 (see section 4.1.2). In both cases, thespectra are due to small-radius, Frenkel excitations.In the case of CuPc/MgF2, it was speculated145 that

Figure 4-26. Photoluminescence peak shift vs 1/d2 forvarious co-(2,6-NV)/SPS and co-(1,4-NV)/SPS assemblies.The shift is measured with respect to the peak position forthick assemblies. Here d is the assembly thickness mea-sured using X-ray reflectivity data (from ref 150).

Figure 4-27. Optical absorption spectra of PTCDI/MgF2multilayer structures with varying PTCDI thicknesses.Also shown is the spectrum from a single, 1000 Å thickPTCDI film. Arrows indicate the prominent excitonic peaksin the spectrum which shift and change oscillator strengthwith PTCDI thickness (from ref 145).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1859

Page 68: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the energy redistribution is due to the presence oftwo different solid phases of CuPc (known as the Rand â forms21) coexisting in the thinnest layers (∼13-26 Å), with both phases contributing their charac-teristic absorption to the total multilayer spectrum.To our knowledge, no spatially extended CT stateshave been identified in PTCBI, and hence no quan-tum size effects are anticipated, consistent withobservations for both of these multilayer systems.Multiple quantum well structures consisting of

alternating layers of CuPc and NTCDA, where thelayer thicknesses were varied from 10 to 50 Å werealso studied.142 A shoulder near the low-energy peakof the CuPc at a wavelength of 700 nm was clearlyobserved in the samples with the thinnest layers andwas less pronounced in the 50 Å thick layers (Figure4-28). Confinement within the organic nanostructuretends to increase the oscillator strength of thisparticular feature, which may result from a changein the energetics provided by the barriers formed bythe adjacent NTCDA layers. Note that feature C,which is due to a CT exciton within the CuPcstack,254,255 is suppressed in the thinnest layers.Quantum size effects have also been suggested to

be the source of shifts in the spectra of ultrathinlayers of Alq3, a material of considerable recentinterest31,32,256 due to its high electroluminescenceefficiency when used in organic light-emitting devices(OLEDs). Ohmori146 observed significant (70 meV)blue shifts in the Alq3 photoluminescence peak as thelayer thickness was decreased from 300 to 20 Å. Inthese experiments, Alq3 was grown in multilayerstacks along with equally thick layers of the widerHOMO-LUMO gap “barrier” material, [N,N′-diphe-nyl-N,N′-bis(3-methylphenyl)-1,1′-biphenyl-4,4′-di-amine], or TPD. Energy barrier arguments146,147were employed to interpret this spectral shift asresulting from confinement of electrons of mass 1.5moin 0.9 eV deep, Alq3 energy wells. A qualitativelysimilar explanation159 was given for shifts observedin both the photoluminescence and absorption spec-tra consisting of alternating organic/inorganic Alq3/MgF2 structures whose individual layer thicknesseswere comparable to those used in the Alq3/TPDstudies.While quantum size effects cannot be ruled out for

Alq3-based structures, they are not likely to exist due

to the Frenkel-like nature of excitons in this mate-rial.257,258 Furthermore, unlike the perylene- andphthalocyanine-based crystalline MQW structures,Alq3 thin films are amorphous. Hence, we expectthere to be very little spatial coherence of thesemolecular excited states, further reducing the pos-sibility that “macroscopic” layer structures exceedinga few Ångstroms in thickness can significantly modifythe exciton binding energy.An alternative explanation for the spectral shifts

is that the Frenkel states bind to the Alq3 surface atthe interface with the adjacent layer. The energyshift in this case arises from the discontinuity inpolarizabilities of the two materials.25,247-249 Sinceexciton energy, to first order, is proportional to 1/ε2,we expect that the exciton energy can be either red-or blue-shifted from its bulk value, depending on therelative dielectric constants of the well and barriermaterials. As discussed in section 4.2, such differ-ences in material polarizability have been taken intoaccount in describing the energy shifts in systemswhere extended excitonic states have been clearlyidentified.176 Indeed, to first order, the polarizationenergy difference simply modifies the well depth bya factor of 1/ε, and hence these two effects can notbe independently determined. Thus, to clearly iden-tify a spectral shift as due to quantum size effects,the spatial dimensions of the exciton of a givenmaterial must first be determined using independentmeasurement methods such as solution or gas phasespectroscopy,197 electroabsorption or electroreflec-tance,177,212,213,255,259 etc.The observations discussed above indicate that

quantum confinement has been observed in systemswhere spatially extended states arise from closeintermolecular stacking within a unit cell. While thisphenomenon is therefore limited to a class of nearlyplanar organic molecules which can be grown intohighly ordered structures, ultrathin layers, neverthe-less, generally have different optical and electronicproperties from those found in the bulk material.These differences can have several sources, rangingfrom material intermixing,144 binding of excitedstates to the numerous heterointerfaces,25,249 dipole-dipole interactions,26,29,183,189,251 chemistry betweencontacting molecular layers, etc. Hence, the realiza-tion of organic/organic and organic/inorganic nano-structures is expected to lead to new physical phe-nomena which can deepen our understanding ofmolecular organic materials, and should also lead toapplications which were not anticipated prior to theability to fabricate this new class of structures. Forexample, several recent theoretical studies of organic/organic and hybrid organic/inorganic nanostructuressuggest that these materials will have unique non-linear optical properties.26-28,218,220,221 In addition,organic multilayers have been found to induce rec-tification of injected charge260 and have been shownto result in output light polarization when incorpo-rated into OLEDs.148,261

One interesting application of organic heterostruc-tures is in the enhancement of the collection ef-ficiency of photogenerated carriers due to dissociationat the numerous, closely spaced heterointerfaces.34A shortcoming of solar cells based on organic materi-

Figure 4-28. Absorption spectra for (a) simple summationof the spectra for 50 Å thick CuPc and NTCDA films; andfor multilayer films with (b) 50 Å per layer and (c), 10 Åper layer. The prominent spectra features are indicated byA, B, and C (from ref 142).

1860 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 69: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

als (see section 5.2) is that the exciton diffusionlength is often significantly less than the opticalabsorption length, leading to low (<1%) power con-version efficiencies.20,190,262 Hence, excitons generatedfar from contacts in organic thin film solar cells oftenrecombine prior to dissociation, and subsequent car-rier collection at the device contacts. While it hasbeen shown172,263 that ordered, OMBD-grown organicfilms can support excitons with very long diffusionlengths (>200 nm) due to the low density of crystal-line defects which act as recombination sites, thephotocarrier collection efficiency is nevertheless low(∼1%).Arbour et al.34 found that organic photoconductors

consisting of alternating multilayer stacks of VOPcand PTCDA, whose individual layer thicknessesranged from 100 to 940 Å, could be used to signifi-cantly enhance the photogenerated carrier collectionefficiency over that obtained using a single layerstructure. Collection efficiencies of three such struc-tures as a function of wavelength are shown in Figure4-29. It was observed that the collection efficiencyin these samples increased linearly with the numberof VOPc/PTCDA interfaces: an effect attributed tothe proximity of interfaces at distances comparableto the exciton diffusion length. As noted above, theheterojunctions in the multilayer photoconductorserve as sites for exciton dissociation into electron-hole pairs which are subsequently separated by thebuilt-in electric fields and collected at the contacts.The enhancement in collection efficiency is quitepronounced, with an increase in quantum efficiencyfrom ∼0.5% for the thickest layers, to 6% for an 8period, 100 Å VOPc/100 Å PTCDA structure. It wasalso found34 that simply codepositing the two molec-ular constituents, VOPc and PTCDA, leads to reducedphotoactivity from that obtained from pristine, single-layer structures. From this, it was concluded thatthe interaction at the heterojunctions of individualmolecules is not responsible for the observed en-hancement. Rather, it is the presence of a structureconsisting of highly pure, ordered molecular layersseparated by molecularly flat, nearly defect-freeboundaries which ultimately improves the carriergeneration and collection efficiencies in this demand-

ing application. The increased fluorescence efficiencyobserved in organic multilayer stacks is also a strongindication of the lack of dissipative recombinationsites at the heterointerfaces.24,34,135,163 To our knowl-edge, the level of material and structural purityrequired to obtain such a pronounced improvementin photoconductor efficiency has only been obtainedusing organic multilayer nanostructures grown byOMBD.Photoconductivity enhancements have also been

observed190,191 in the hybrid organic/inorganic mul-tilayer system, CuPc/TiOx. It was suggested that thelarger electron affinity of TiOx (4.2 eV) as comparedwith CuPc (3.1 eV) acts to collect and conduct chargesin the TiOx layers which were originally photogener-ated in the adjacent CuPc layer. In these experi-ments, multilayers consisting of 50-400 Å thickCuPc layers were grown between TiOx layers ofcomparable thickness. The growth temperature wasmaintained at 200 °C in order to obtain R-CuPcwithout significant interface roughening. Takada etal.191 observed that the photoconductivity of themultilayers at high incident optical flux (∼100 mW/cm2) was 40 times larger than that obtained for CuPcsingle layer samples, with the 400 Å thick CuPcsamples exhibiting the largest photoconductivities.This increase was attributed to a two-step process,beginning with exciton generation in the CuPc uponabsorption of the incident light. Next, the excitonseither dissociate at O2

- trap centers induced ongrowth of the TiOx films (which occurs in the presenceof an oxygen plasma), or they diffuse to a nearbyCuPc/TiOx heterointerface where dissociation canalso be promoted. Once the free electrons are pro-duced, they drift into the high electron affinity,relatively high mobility TiOx and then to the contactsat the device periphery. The decrease in carriercollection efficiency at lower incident optical fluxesis possibly due to the presence of interface defectswhich act as recombination centers for both excitonsand free charges.

4.4. Implications and Applications of ExcitonConfinement in OMCsWe have shown that accurate methods to analyti-

cally treat CT excitons in tightly packed neutralmolecular crystals have been developed to describeseveral different spectroscopic features of vacuum-deposited organic nanostructures. It was found thatclosely packed organic crystals are characterized bya large overlap between the orbitals in adjacentmolecules, resulting in delocalization of the CTcharge pair. In this case, the crystalline pseudopo-tential (or the on-site correlation energy) becomescomparable to or even larger than the energy band-width, with the consequence that the CT exciton canbe represented by a hydrogen-like wave function.Using a Hamiltonian modified to account for the

calculated molecular polarization and suitable hy-drogenic trial wave functions, the electroabsorptiondata for homogeneous PTCDA thin films, as well asthe absorption spectral shifts observed in PTCDA-based multiple quantum wells was accurately fit,further testing the picture of an extended CT statein certain OMCs. The analyses of such experimentsgive independent and consistent estimations of the

Figure 4-29. Photocurrent yield for various VOPc/PTCDAmultilayer assemblies. The subscript after the layer de-scription indicates the number of bilayers in the particularassembly. A significant enhancement in photocurrent yieldis observed as the layer thickness is increased (from ref34).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1861

Page 70: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

exciton radii for PTCDA along different crystal axes.This treatment was extended to other exciton sys-tems found in a wide range of interesting neutralorganic molecular and inorganic semiconductors (suchas GaAs), suggesting that Wannier and CT excitonshave a common physical origin which can be ap-proximated by a hydrogenic wave function. Thisprovides conclusive evidence that PTCDA and simi-larly close packed molecular compounds such as192,193BTQBT and possibly even C60 and its analogs259 forman important “transitional” class of materials thatshare many properties with both conventional semi-conductors and insulator-like organic molecular crys-tals.While these results can be easily generalized to

explain a variety of observations of the opticalproperties relating to the layer thickness or aggregatesize of both organic and inorganic nanostructures,they also provide a new viewpoint with which toconsider vdW-bonded solids represented by the ar-chetype compound: PTCDA. The broad and richrange of physics developed for understanding quan-tum confinement in inorganic semiconductors whichhas ultimately led to the precise control and manipu-lation of their densities of states to attain desiredoptoelectronic properties, appears to be equally ap-plicable to organic nanostructures provided there issufficient charge delocalization resulting from detailsof the crystal structure. In that sense, “transitional”materials such as PTCDA and related compoundshave bridged the gap between highly delocalizedsemiconducting systems and insulators.Indeed, it has already been found that multilayer

structures exhibit new properties beyond those ex-hibited in conventional, homogeneous thin films.Multiple phases and mixtures of compounds havebeen identified for both organic/organic and organic/inorganic multilayers with interesting optical proper-ties. In two cases,34,191 an increase in photoconduc-tivity of the multilayers as compared with single filmstructures has been found, suggesting that thesetypes of engineered materials may eventually findapplication in solar energy conversion. The observa-tion of quantum confinement in MQWs consisting ofclosely packed OMCs24 also suggests that thesestructures can be used to tailor the electronic densi-ties of states of the constituent materials. Followingthe example of inorganic semiconductor structures,the density of states in the reduced dimensionalityof MQWs is useful for, among other applications,decreasing laser threshold current.12 Hence, onepotential application of organic MQWs may be in therealization of organic solid-state current injectionlasers with reduced threshold currents (and thusreduced power dissipation and longer operationallifetime).These represent only a few of the interesting

physical and technological possibilities opened up bythe recent achievement of both organic and organic/inorganic (hybrid) multilayer nanostructures in sev-eral laboratories worldwide. In the following section,we will briefly discuss a few practical optoelectronicdevices which have recently been demonstrated basedon high-purity, vacuum-deposited organic thin films.

5.0. Applications of Thin-Film Structures Grownin VacuumA primary motivation for the extensive research

over the last several years concentrating on thegrowth and physics of vacuum-deposited organic thinfilms is their very real potential for use in applica-tions which are not accessible to more conventional,inorganic semiconductors. Indeed, the recent dem-onstration of efficient electroluminescence from or-ganic thin-film devices31,32,264 promises to transformthe flat panel display industry, with the potential ofreplacing liquid crystal displays with an entirely newgeneration of efficient, emissive, full-color flat panelsbased on light-emitting organic devices. In morerecent developments, organic thin films are showingpromise for use as thin-film transistors265-270 whichmight eventually replace amorphous or polysiliconTFTs currently used in the back planes of active-matrix liquid crystal displays (AMLCD). These newdevelopments must also be placed in the context oflong efforts and progress which has been directed atemploying organic thin films for solar energy conver-sion20,271 and in sensors of various kinds. Finally,vacuum-deposited OMCs have also been proposed asmaterials with large second- or third-order opticalnonlinearities,18,26-28,160,218,220,221,272-279 or large ø(3).There are, in addition, many, somewhat less conven-tional uses of organic thin films deposited in vacuum,including waveguides and optical couplers,22,102,280,281organic/inorganic photodetectors,38,282 lasers, etc.The primary attraction of OMCs is their potential

low cost and the extreme flexibility which the deviceengineer has in choosing a material whose propertieshave been specifically tailored to meet the needs ofa particular application. The materials are easilyintegrated with conventional semiconductor devices,thereby providing additional functionality to existingphotonic circuits and components. This potential,however, must be balanced against the problemswhich have traditionally impeded the acceptance ofOMCs for use in active electronic or optoelectronicdevice applications. These problems include unstabledevice characteristics, sensitivity to adverse environ-ments (e.g., temperature, humidity, oxygen, etc.),nonideal metal/organic contacts, and lack of repro-ducibility of material composition, purity, and fabri-cation conditions. From the foregoing discussion onOMBD of molecular organic thin films, it is clear thatthe ultrahigh-vacuum environment characteristic ofthis deposition process can provide the necessarymaterial purity and structural and chemical repro-ducibility necessary in modern, high-performanceoptoelectronic device applications. While the costsand complexities associated with UHV depositionprocesses may offset the attractive (but possiblymisleading) “simplicity” often attributed to organic-based devices, it is clear that the advantages inperformance of such structures outweigh these ap-parent disadvantages. Furthermore, while purityand structural precision are key to the ultimatesuccess of all optoelectronic device technologies, it isnot clear how sophisticated the OMBD system mustbe to achieve acceptable device performance. At thispoint, OMBD serves as our most powerful tool forinvestigating the detailed growth and physical char-

1862 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 71: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

acteristics of OMCs, and hence will ultimately be ableto address questions regarding the need for UHV inthe production of practical display, transistor, NLO,or other molecular organic thin-film device applica-tions.In this section, we briefly review several potentially

important applications where OMCs have been usedto improve upon or expand the functionality of thedevice beyond that which can be achieved usingconventional, inorganic semiconductors. Indeed,rather than focusing on all the many areas whereOMCs have proven to have properties amenable topractical application, we concentrate here on thoseapplications where the use of OMCs may transformcurrent device technology by providing a significantadvantage over present methods. This section willtherefore describe the use of OMCs in organic/inorganic (OI) semiconductor heterojunction devices,solar cells, organic light-emitting devices (OLEDs)and lasers, and thin-film transistors (TFTs). Thediscussion is by no means intended to be encyclope-dic, but rather should provide a useful introductionto the state-of-the-art of some of the most importantcurrent and future devices based on vacuum-depos-ited organic molecular thin-film materials.

5.1. Integrated Organic-on-InorganicHeterojunction DevicesThe first demonstration38 of an organic/inorganic

semiconductor device employing a vacuum depositedthin film consisted of a 1000-2000 Å thick PTCDAfilm on (100) p-Si (see inset, Figure 3-52). Thecontact between the organic and inorganic layersforms a rectifying heterojunction (or organic-inor-ganic heterojunction, OI-HJ). Given the p-type na-ture of PTCDA, when deposited on a p-type semi-conductor substrate an isotype heterojunction isformed, whereas an anisotype heterojunction is formedon an n-type substrate. These OI-HJ devices havebeen shown to be useful in the nondestructive map-ping of the free-carrier concentrations in the inor-ganic substrates. Indeed, rectifying OI-HJ contactshave been successfully demonstrated for this purposeon a wide range of group II-VI, III-V, and IVsemiconductors283-286 using a variety of organic thin-film materials.37,38,170,283,287,288 They have a very highreverse breakdown voltage and low dark currentsdetermined by the properties of the inorganic sub-strate rather than the organic thin film itself. Thisallows for depletion of the free carrier distributiondeep into the substrate by application of large reversevoltages. Employing conventional capacitance-volt-age profiling of the free carrier concentration in thesubstrate, this parameter is obtained as a functionof depth into the semiconductor.63,285,289,290 Once thefree carrier concentration is obtained, the organicthin film and contacts are nondestructively removedfrom the semiconductor surface (e.g., by dissolvingPTCDA in a high pH solution), allowing for furtheruse and processing of the characterized material.Furthermore, the nondestructive deposition of theorganic film onto inorganic substrates has beenshown to be a means for quantitatively determiningthe electronic surface state density distribution of thesubstrate.170,291-293 Surface states are known togovern the properties of many semiconductor devices

such as metal-insulator-semiconductor (MIS) de-vices and surface-passivated rectifiers and transis-tors. Hence, obtaining detailed characteristics of thesemiconductor surface is a crucial first step in theprocess of fabricating high-performance devices. Thenoninvasive organic/inorganic HJ contact has there-fore proven to be an extremely useful new tool inobtaining the characteristics of electrically activedefects at semiconductor surfaces, which have oth-erwise not been possible to obtain using conventionalanalysis techniques which tend to be more destruc-tive of the fragile, as grown surface properties.The OI-HJ can be understood in terms of the

energy band diagram63 in Figure 5-1. In this case,PTCDA is in contact with an n-type semiconductorsubstrate. This diagram illustrates an “optical band-gap” between the HOMO and LUMO levels of 2.2 eV,along with the relatively smaller bandwidth charac-teristic of the organic material. The heterojunctionenergy offset, φB, controls the injection of charge fromthe metal/organic contact (which is found to be nearlyohmic for many metal/organic systems such as In/PTCDA) into the semiconductor substrate. Alsoshown is an interface region of thickness δss whichis due to the thin native oxide that typically existsat the semiconductor surface. This interface regionis a source of charge of density Dss due to danglingbonds at the semiconductor surface (see Figure 3-50).These surface states can exist at the as-grown (orintrinsic) semiconductor surface, as well as at sur-faces exposed to atmosphere, as is typically the casefor OI-HJ devices. The presence of the energybarrier, φB, results in the rectification of current bythe OI-HJ. In reverse bias (i.e., where the substrateis at a positive potential relative to the metal contact),the carriers must overcome the barrier potential.Under forward bias, the carriers are injected fromthe substrate into the organic thin film, whichultimately limits the current due to the small mobili-ties (µ < 1 cm2/V-s) typical of ordered molecularmaterials such as PTCDA. The current-voltagecharacteristics of the OI-HJ are described by eq

Figure 5-1. Proposed energy band diagram of an OI-HJconsisting of PTCDA on an inorganic semiconductor sub-strate. In this diagram, εg is the semiconductor energy gap,φB is the heterojunction LUMO/conduction band barrierpotential, and δss is the thickness of a layer at thesemiconductor surface. Inset: Proposed densities of statesat the OI-HJ interface (from ref 63, p 199).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1863

Page 72: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(3.14), where the saturation current, Jso is determinedby the barrier height171 viz.

Here, Ncs is the semiconductor effective conductionband density of states, ⟨vc⟩ is the mean carrier velocityin the organic semiconductor (∼500-1000 cm/s forPTCDA), and vd is the diffusion velocity in theinorganic substrate. The barrier height can beinferred from the current-voltage characteristics ofthe OI-HJ device, or can be measured directly usinginternal photoemission.217 Barrier heights (φB) havebeen obtained for PTCDA on various p-type groupIV and III-V semiconductor substrates171 and arelisted in Table 9. From φBp, we can derive the energyoffset (∆Ev) between the HOMO and the valence bandof the semiconductor using ∆Ev ) qφBp + ∆s - ∆o,where ∆s and ∆o are the energy differences betweenthe equilibrium Fermi levels in the bulk of theinorganic semiconductor and the organic thin film,respectively. Values of∆Ev for several PTCDA/p-typeinorganic semiconductor systems are also listed inTable 9.It has been found that the characteristics of OI-

HJ diodes follow the ideal expression in eq 3.14,exhibiting a nearly exponential increase in currentwith voltage in the forward biased direction (Figure5-2) until the mobility-limited internal organic filmresistance results in a decrease in current at highvoltages.171 In contrast, the reverse-biased charac-teristics (Figure 5-3) are limited by dark currentgeneration in the inorganic semiconductor, until theelectric field at the OI-HJ becomes sufficiently largeto induce avalanche breakdown in the substrate.37Other, less ideal sources of dark current and highvoltage breakdown can also be observed, dependingon the characteristics and quality of the substratematerial. However, when the applied electric fieldis sufficiently high to induce avalanche, such as isthe case for the PTCDA/p-Si device in Figure 5-3,large photocurrent multiplication can be observed.In the device shown, photogenerated electron mul-tiplication in the p-Si substrate exceeds 100.These results indicate that OI-HJ devices can be

used as photodetectors which are both simple tofabricate and have performance characteristics com-parable to or better than that obtained using con-ventional inorganic semiconductor devices. An in-

teresting example of this application is an PTCDA/p-Si photodetector where the metal top contact wasreplaced by a sputter-deposited, transparent indium-tin-oxide (ITO) hole injecting contact.282 This devicecould be reverse biased into avalanche breakdown,exhibiting uniform gain characteristics in the pres-ence of very low primary dark currents (6 µA/cm2)even at high reverse voltages (>50 V). The band-width of the device was ultimately limited by the hole

Table 9. Barrier Energies for PTCDA-on-InorganicSemiconductor Diodes (From Ref 171)

inorganicsubstrate

majoritycarrier type qφB (eV) ∆Ev (eV)a

Si n 0.39 ( 0.01p 0.56 ( 0.02 0.50 ( 0.02

Ge p 0.30 ( 0.03 0.25 ( 0.04GaSb p 0.37 0.29 ( 0.04

n ohmicGaAs n 0.39

p 0.52 0.43 ( 0.05InP n 0.30

a ∆Ev for p-type substrates only. For these calculationsm*) mo for PTCDA. The value for p-Si is measured, whereasfor other materials, ∆Ev is calculated using φBp obtained fromref 284 and reproduced here. Doping of p-Si and p-GaSb: 5× 1015 cm-3; p-Ge: 5 × 1014 cm-3; and p-GaAs: 2 × 1016 cm-3.

Figure 5-2. (a) Forward current-voltage characteristicsof an In/2000 Å PTCDA/p-Si OI-HJ device measured atseveral different temperatures (from ref 65). (b) Saturationcurrent density vs 1/T for the diode in a. The solid linerepresents a best fit to the data, giving a barrier diffusionpotential of φBp ) 0.56 V (from ref 171).

Figure 5-3. Reverse current vs voltage characteristics ofIn/PTCDA/p-Si devices obtained at room temperature.PTCDA thicknesses of d ) 100 and 2000 Å were depositedon 10 Ω cm p-Si substrates. Also shown is the photocurrentmultiplication factor for the devices (from ref 37).

Jso ) [qNcs⟨vc⟩/(1 + ⟨vc⟩/vd)] exp(-qφB/kT) (5.1)

1864 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 73: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

transit time across the PTCDA layer. By making thislayer very thin (∼100 Å), bandwidths exceeding 500MHz are possible using such structures.To be practical, it is essential that OI-HJ diodes

be stable over long periods when exposed to adverseenvironmental conditions such as elevated temper-atures and high humidity. Prabhakar et al.294 dem-onstrated both photolithographic patterning andenvironmental passivation of an OI-HJ diode usinga relatively low temperature (125 °C) pyrolytic pro-cess to deposit a SiO2 encapsulating layer onto theorganic surface. The dark current characteristicsunder reverse bias of the resulting diodes weremeasured as a function of time and temperature. Thedevices were found to be stable up to temperaturesas high as 100 °C over extended periods (>3000 h).These results suggest that molecular organic materi-als can be patterned and passivated using photo-lithographic techniques adapted for this purpose andthat many organic materials have physical charac-teristics which are compatible with applicationswhere elevated temperatures are required.295Indeed, devices with very small patterns such as

optical modulators296 and narrow gate (∼2 µm) OI-HJ transistors based on OMC/InP contacts297 havebeen demonstrated using patterning techniques de-veloped specifically for their compatibility with thechemistry and physical properties of molecular or-ganic materials. A cross section of the transistorstructure and its drain current characteristics areshown in Figure 5-4. The problem of micropatterningof the 1.5 µm gate contact was solved by creating an

undercut pattern in the substrate, which provides fora “self-aligned” structure. The fabrication proceededas follows:297 After ion implantation of the semiin-sulating (100) InP substrate to form the channel andcontact regions, a Si3N4 mask was deposited byplasma-enhanced chemical vapor deposition (PECVD).This mask was photolithographically patterned toopen areas for the gate, source and drain contacts.After standard metal deposition of the source anddrain, an etchant was used which selectively exposesthe (111) planes of InP tilted at 54° to the substratenormal while undercutting this pattern beneath theSi3N4 mask. This undercut etched groove providesthe pattern for the organic thin film deposition. Inthis case, N,N′-dimethyl-3,4,9,10-perylenetetracar-boxylic diimide (DM-PTCDI) was deposited at anangle such that it covered regions underneath theoverhang. Finally, the gate metal (In) was depositedwith the wafer positioned perpendicular to the metalbeam flux to ensure the metal did not short to thesubstrate. These narrow gate transistors had excel-lent output characteristics as shown in Figure 5-4b.For example, the transconductance was gm ) 80 mS/mm, comparable to that obtained with conventionalInP FETs. The bandwidth of such OI-HJ devices isultimately limited by the carrier transit time acrossthe thin organic layer. In practice, this can exceed1 GHz for 100 Å thick, ordered films.282 Moresignificantly, the devices showed no indication ofdrain-current instability (or “drift”) commonly ob-served in InP metal-insulator-semiconductor de-vices.298 The organic heterojunction also enhancedthe gate barrier height (0.54 eV) over most metal-semiconductor contacts (where the barrier height ofmost metals to INP is only ∼0.4 eV), thereby greatlyreducing gate leakage. In effect, the use of OI-HJgates significantly improved the device performanceover that which could be obtained using conventional,all-inorganic materials systems.It was noted above that one of the attractive

aspects of molecular organic thin films is their abilityto be nondestructively integrated with inorganicsemiconductor devices. One example device, dis-cussed in section 3.3.2.4, is a recently demonstratedmolecular organic waveguide coupled to anIn0.53Ga0.47As photodiode for use in optical com-munications systems,22 and shown schematically inFigure 5-5. In this device, a PTCDA waveguide isgrown on a previously vacuum-deposited and pat-terned amorphous Teflon ridge on the (100) surfaceof an InP substrate held at Tsub ) -100 °C duringgrowth. Amorphous Teflon was employed as thewaveguide buffer since its index of refraction (n )1.15-1.30) is lower than the PTCDA index eitherperpendicular (n⊥ ) 1.36) or parallel (n| ) 2.12) tothe substrate,175 allowing for propagation of lightindependent of polarization direction. The waveguidewidth was varied between 4 and 20 µm. A second,vacuum-deposited Teflon buffer layer was depositedonto the PTCDA surface, thus completing thewaveguide. The waveguide was butt-coupled into theIn0.53Ga0.47As optical absorption region in a reverse-based, planar p-i-n photodiode with sensitivityextending to a wavelength of λ ) 1.65 µm. Waveguidelosses between 2.7 and 4.6 dB/cm were measured,

Figure 5-4. (a) Cross-sectional diagram of a self-alignedorganic-on-inorganic semiconductor transistor (OISTR)employing a 500 Å thick layer of DM-PTCDI to form acontact barrier to the ion implanted, InP substrate. (b)Drain current characteristics of the OISTR (horizontalscale, 1 V/div; vertical scale, 5 mA/div; gate voltage step,0.5 V) (from ref 297).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1865

Page 74: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

with the higher losses corresponding to the narrowestguides due to optical out-coupling at the roughwaveguide edges.The most interesting feature of these guides is that

they were polarization preserving.22,115 That is, ifpolarized light is launched at the waveguide input,the output polarization after propagation along a 0.5cm long guide remains in the same direction as atthe input. Since the optic axis (indicated as the “c”axis in Figure 5-5) of the OMC has a unique orienta-tion with respect to the PTCDA unit cell, this lack ofpolarization rotation indicates that the c axis lies inthe waveguide x-z plane. This, coupled to therelatively low loss of the waveguides, suggests thatthe growth and fabrication conditions used in pro-ducing these integrated devices result in molecularalignment and smooth surface morphologies overextended (∼1 cm) distances for QE-grown films. Thisunique aspect of OMCs grown by OMBD can lead tonovel applications such as the edge-coupled inte-grated waveguide/photodetector shown in Figure 5-5.

5.2. Organic Solar CellsIt has long been recognized that thin-film OMCs

have promise for solar energy conversion due to thepotential for very low cost manufacture of solar cellsemploying such materials.20 In particular, crystallineorganic semiconductors such as the Pc’s, perylenesand other relatively low molecular weight polyaceneshave been identified as having the greatest potentialfor this purpose due to their ability to be controllablydeposited via OMBD and other more conventionalvacuum techniques, their relatively high purity,environmental robustness, and high mobilities (hencerelatively low series resistance).19,271,299-310 Morerecently, solar cells employing these materials incombination with C60 have also been explored311,312for many of the same reasons. In spite of consider-able research in both bilayer organic p-n junctionand Schottky barrier crystalline organic photovoltaic(PV) cells over the past 20 years, however, there hasyet to be a demonstration of a cell whose character-istics are adequate for even the most undemandingof solar conversion applications.20 Their poor per-formance can be ascribed principally to the followingcauses: (i) Due to the intrinsic nature of photocon-ductivity in crystalline organic materials where freeelectron-hole pairs are generated in a second-orderprocess following absorption and exciton generation,

the efficiencies realized to date are low (typically<1%for AM0 to AM2 illumination). (ii) Due to the limitedπ-orbital overlap between adjacent molecules in anorganic crystalline stack and due to the numerousdefects in the crystalline order in the depositedmaterials, the free carrier mobilities in many organ-ics are low (typically <10-3 cm2/(V s)). This leads tohigh film resistance, and hence low power conversionefficiency. (iii) The exciton diffusion length (LD) istypically considerably shorter than the optical ab-sorption length, 1/R, where R is the absorptioncoefficient (see section 4.3). Hence, the exciton isoften photogenerated far from an interface wheredissociation into free carriers can occur, leading torecombination and a concomitant loss in quantumyield. (iv) A final consideration is one of environ-mental stability of the organic thin-film materials,and in particular the stability of the intramolecularbonds to prolonged exposure to ultraviolet radiation.While these problems currently prevent the real-

ization of fully organic, low-cost PV cells, significantadvances in OMBD growth and processing of molec-ular organic thin-film devices made by researchersworldwide suggest that such materials may eventu-ally find practical application for low-cost solarenergy conversion. As discussed in section 4.3, MQWstructures can significantly increase the quantumyield and carrier diffusion lengths in organic materi-als. Indeed, extending the photoactive region widthby inserting multiple junctions, each of thickness∼1/R, represents a recent, promising approach madepractical due to the uniformity control inherent inOMBD growth.34 Furthermore, the same films grownwith a high degree of stacking order (e.g., those basedon perylenes and Pc’s), typically also have enhancedπ-orbital overlap, thereby increasing carrier mobilitywith its concomitant decrease in bulk layer resis-tance.There are two basic solar cell structures: the metal/

organic Schottky-type cell, and the fully organicbilayer cell which employs two ohmic metal contacts.The Schottky cell consists of a light-absorbing organicfilm sandwiched between two metal contacts: onerectifying and the other ohmic. For a p-conductingorganic material, the Schottky contact consists of alow work function metal such as Al, Mg, or In,whereas for n-conductivity films, a high work func-tion metal such as Au is preferred. Charge separa-tion is induced by exciton dissociation in the space-

Figure 5-5. Schematic view of an integrated organic waveguide/photodetector with axes indicating the relationship ofthe PTCDA optic c axis relative to the waveguide (from ref 22).

1866 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 75: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

charge region due to the built-in electric field at therectifying contact, which therefore requires thatoptical absorption occur in this “photoactive” regionas well. A photoactive region width of ∼200-400 Åis typical.In the bilayer cell, charge separation occurs at the

organic heterojunction whose built-in potential isdetermined by the size of the HOMO-LUMO gapenergy difference300 between the contacting materi-als. The HOMO offset is equal to the difference ofionization potentials (IP) of the two contacting ma-terials, and the LUMO offset is equal to the HOMOoffset plus the difference in HOMO-LUMO gapenergies (EH-L) of the two materials. To avoidcreating barriers at the organic heterojunction whichimpede carrier transport, it is necessary for thematerial with the larger hole mobility to also have alarger IP and IP - EH-L (i.e., the electron affinity)than the high electron mobility material. Absorptioncan occur in either of the two organic films, therebydoubling the photoactive region width where opticalabsorption can result in efficient charge separationby the built-in field. Note that it is not importantwhether a “p-n”-like junction or a simple isotype (i.e.,p-P or n-N) heterojunction is employed, since it isonly the diffusion potential created at the hetero-junction due to the HOMO-LUMO gap offsets thatis effective in carrier drift.Whether or not an organic material shows p-type

or n-type character depends only on the backgroundfree carrier concentration of the material, which isprimarily a function of unintentional doping bydefects or impurities.268,300,313 These impurities thendetermine the position of the Fermi energy withinthe HOMO-LUMO gap. A much more significantfeature of an organic semiconductor (which is inde-pendent of Fermi energy depth in the gap) is whetherelectrons or holes have a higher mobility. Unlike thefree carrier concentration, this is an intrinsic featureof the organic material, determined by the crystalsymmetry and periodicity which provide a higheroverlap of HOMO (LUMO) levels in adjacent mol-ecules, giving rise to a higher hole (electron) mobil-ity.80 Indeed, sometimes the chemical character ofthe molecule itself is not the most important consid-eration in determining this overlap. For example, theanhydride groups in PTCDA have a strongly electrondonor character. However, molecular orbital calcula-tions of the molecule show that the frontier orbitalsfor this model compound are on the perylene core,while the anhydride orbitals are located deep in theLUMO band, thereby contributing little to the elec-tronic conduction properties of the crystalline PTCDAfilm.43,58 Current conduction experiments indicatethat the hole mobility of PTCDA exceeds that of theelectron mobility by several orders of magnitude,suggesting the importance of band structure inultimately determining the carrier conduction prop-erties of a material.172,173,224 This distinction betweenmajority carrier type and mobility is very oftenmisunderstood by the organic thin-film device com-munity. Hence, the assignment in the literature ofa material as “n” or “p” type to explain solar cellbehavior must be viewed with caution. It has beenshown that isotype heterojunctions formed betweenmaterials of the same conductivity type can equallywell lead to excellent diode properties, and can be

used to fully quantitatively explain the forward andreverse biased characteristics of such rectifying junc-tions.300,314,315 Hence, the most reliable data availableconcern materials where mobility and conductivitymeasurements have been independently measuredusing such techniques as the analysis of metal-organic Schottky barrier conduction, and time offlight measurements.37Low quantum yield (η) in organic PV cells is

attributed to the intrinsic nature of the photocon-ductivity process, illustrated in Figure 5-6. Theabsorbed solar energy first results in the efficientgeneration of excitons in the film bulk316 (describedby the process So + hν w So*, where So and So* denotethe molecular ground and excited states, respec-tively). Subsequently, the exciton drifts to a contact,impurity, interface or other inhomogenity within thestructure (denoted M), at which point it dissociatesinto a free carrier pair (via So* + M w e- + h+), oralternatively, suffers recombination. The recombina-tion process accounts for a loss of roughly 95% to 99%of the excitons. Although exciton ionization is notfully understood,226 it is often attributed to dissocia-tion in the presence of the built-in electric fieldsurrounding the crystal defect. If this dissociationoccurs in an otherwise neutral region of the film, thephotogenerated free electron-hole pair may remainlocalized within the vicinity of the defect where theyrecombine without being collected in the externalcircuit. Hence, to significantly increase η, two condi-tions must be met: (i) The film must be relativelyfree of defects which generate local electric fields, and(ii) the exciton and free carrier diffusion lengths mustbe sufficiently long such that these particles canmigrate to regions of the film where the built-inelectric field from an adjacent junction is high enoughto separate the free electron-hole pairs prior torecombination.If dissociation occurs at a contact or organic het-

erojunction, the built-in potential at that junctionleads to charge separation, followed by carrier trans-port to the contacts and delivery of power to anexternal load. This two-step mechanism (exciton

Figure 5-6. Schematic diagram of the photoconductivityprocess in organic thin film solar cells.

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1867

Page 76: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

generation followed by carrier generation) is typicallya very low efficiency process which is not expectedto lead to power conversion efficiencies exceeding∼5%. To date, the highest power conversion ef-ficiency (1.8%) has been reported for a PTCDA/CuPcbilayer cell174 under an illumination intensity of 10mW/cm2. Here, the power conversion efficiency isdefined as

where ff is the forward-biased “fill factor”,19 Voc is theopen circuit voltage, Isc is the short circuit current,and Pinc is the incident optical power. These vari-ables are defined in Figure 5-7. At the higherincident power of 75 mW/cm2 (equivalent to AM2illumination), a maximum power conversion effi-ciency of ηP ) 0.95% was reported for a CuPc/PTCBIbilayer cell.271 One problem with organic-based pho-tovoltaic cells is their high contact and series resis-tances which lead to space-charge build-up withincreasing incident power. It has been shown307 thatthere is a roughly 2-fold decrease in efficiency result-ing from space charge effects as the power densityincreases from 10 to 75 mW/cm2, suggesting that theresults for the CuPc/PTCBI and PTCDA/CuPc cellsare nominally equal. The I-V characteristics of

these p-P heterojunction cells,300 as well as schematiccross sections of both the PTCDA/CuPc and CuPc/PTCBI cells are shown in Figure 5-7, parts a and b,respectively. Of particular note is the low seriesresistance of the latter structure, as indicated by theabrupt increase in current at V > Voc. This leads tothe high fill factor observed, although the authorsfound that the ff characteristics degrade after pro-longed exposure of the cell to air.271

A summary of recent results for organic PV cellsusing both Schottky and bilayer geometries is pro-vided in Table 10. The best results obtained afterthose in Figure 5-7 were reported for a H2Pc/DM-PTCDI bilayer cell with a built-in potential enhancedby H2 doping, due to a lowering of the Fermi levelwithin the DM-PTCDI.303 Increasing the built-inpotential (and hence increasing Voc) has been at-tempted using several different approaches, the mostnotable being insertion of additional organic layers304or even fabricating series-connected “tandem” PVcells in a single stack.302 These methods have provensuccessful in increasing Voc which extends the effec-tive width of the photoactive region, although theygenerally also result in an increase in series resis-tance (with its concomitant decrease in ff), therebylimiting the overall power efficiency of the devicesto values somewhat less than those reported for thebest, single junction cells.

It has been suggested that increased order withinthe film also results in an increased exciton andcarrier diffusion length, which ultimately shouldproduce a higher carrier mobility along the molecularstacks. These effects have been extensively stud-ied172,305 in the model compound, PTCDA, which iswell-known to have extremely close π-π stacking andlong-range stacking order. Assuming that all excitondissociation (and subsequent free carrier generation)occurs at contacts in a metal/organic/metal cell,Ghosh and Feng262 have shown that the photocurrentdensity (J) is governed by exciton diffusion, J ) -Dex

dn(x)/dx, where Dex is the exciton diffusion constant,and n(x) is the exciton density as a function ofdistance with respect to the contact on which thelight is incident (i.e., the “near” contact), or thecounter electrode (the “far” contact) depending onwhere the exciton dissociation occurs. Now, n(x) issimply proportional to the optical flux (φ(x)) at x,given by φ(x) ) φo exp[-Rx], where φo is the flux atthe near contact. From this analysis, it can be shownthat the quantum yield due to exciton generationprimarily at the near and far interfaces follows astraight line when 1/ηNEAR or 1/[ηFAR exp(Rd)] areplotted as a function of 1/R, respectively. Here, d isthe film thickness. Results of this analysis for threedifferent wavelength regions for PTCDA films grownby OMBD172 are shown in Figure 5-8. At shortwavelengths where Frenkel excitons are generated(c.f. Figure 4-2b), the diffusion length is LD ) 88 ( 6nm, consistent with independent studies of Karl etal.305 However, the diffusion length of the CT excitonat low energies (∼2.0 eV) is LD ) 225 ( 15 nm which,to our knowledge is the longest diffusion length

Figure 5-7. Two representative current vs voltage char-acteristics for the highest efficiency cells reported to date.(a) PTCDA/CuPc bilayer heterojunction cell with a powerconversion efficiency of 1.8% at an optical power of 10 mW/cm2 (from ref 174). (b) PTCBI/CuPc cell with an efficiencyof 0.95% at 75 mW/cm2 (from ref 271). Both light-on(dashed lines) and light-off (solid lines) characteristics areshown. Also, various terms used to describe solar cellperformance are indicated in b: Voc ) open circuit voltage,Isc ) short circuit current. The maximum power rectangleshown by the cross-hatched region has an area equal toImaxVmax. The cell fill factor is defined by ff ) ImaxVmax/IscVoc< 1.

ηP ) ffVocIsc/Pinc (5.2)

1868 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 77: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

measured for an exciton in an organic thin film. Thevoltage dependence of the quantum yield apparentin the figure arises from improved carrier collectionefficiency at either the near or far contact (dependingon the sign and magnitude of the voltage) due toincreased fields within the film bulk. The longdiffusion lengths observed for OMBD grown samplesclearly suggests the need for extended stacking orderto achieve long diffusion lengths with its concomitantincrease in quantum yield.Finally, a detailed study of the effects on perfor-

mance due to material purity and deposition condi-tions on bilayer DM-PTCDI cells employing phtha-ocyanine, perylene and porphyrinic compounds as thesecond organic material, have also been recentlyreported.307 For example, the source materials forZnPc/DM-PTCDI cells were purified prior to deposi-tion using three different methods: solvent extrac-tion, reprecipitation of ZnPc from concentrated sul-furic acid, and multiple step sublimation in vacuumas discussed in section 2. The primary figure of meritwas ηP, which was highest for sublimed materials(with ηP ) 0.29%) and lowest (by a factor of 3) formaterial which underwent solvent extraction (ηP )0.07%). From these results, it was inferred thatmaterial purity is a key parameter in achieving highefficiency solar energy conversion. Impurities acting

as ionic trapping centers, or which disrupt thestacking order of the films, ultimately increase seriesresistance, charging (and hence space-charge build-up), and decreases the mobility of the charge carriersand excitons in the films. Furthermore, trappedcharge can shift the Fermi-energy positions withinthe films, thereby decreasing (or increasing) Vbi,ultimately determining the width of the photoactiveregion.The study was extended to examine the effects of

deposition rate, which was varied between 0.1 and15 Å/s. For these materials, no systematic surfacemorphology dependence was observed over this rangeof rates, suggesting that the film order was alsounaffected. This was also consistent with the findingthat ηP and other device performance parameterswere not dependent on deposition rate.An additional parameter which has been system-

atically varied in the growth of DM-PTCDI/ZnPc PVcells is substrate temperature,307 varied from 25 to200 °C. Due to roughening of the surface at thehighest temperatures (>100 °C), short circuits de-veloped and the PV cells ceased to function. Belowthis temperature, there is a systematic decrease inηP from 0.33% at Tsub ) 25 °C to 0.28% at Tsub ) 100°C. This reduction in ηP was attributed to elimina-tion of shorts (which reduce Voc) as the growth

Table 10. Representative Molecular Organic-Based PV Cell Results

cell composition Voc (mV) Isc (mA/cm2) Pinc (mW/cm2) ff ηP (%) comments ref(s)

Ag/PTCBI/CuPc/ITO 450 2.3 75 0.65 0.95 271

In/PTCDA/CuPc/ITO 260 2.0 10 0.35 1.8 low Pinc 300

Au/H2Pc/DM-PTCDI/ITO 550 2.6 100 0.30 0.77 H2 doped 303580 2.0 0.23 0.49 NH3 doped660 1.6 0.22 0.41 undoped

Au/H2Pc/DM-PTCDI/PTCBI/ITO 370 0.5 25 0.27 0.20 304

Au/H2Pc/PTCBI/ITO 370 0.18 25 0.32 0.08 304

Al/merocyanine/Ag 1200 0.18 78 0.25 0.62 Schottky cell, best at d ) 150 Å 262

Al/ZnPc/Au 590 5.6 × 10-4 0.1 0.1 3 × 10-4 Schottky cell 306

Au/ZnPc/DM-PTCDI/ITOa 450 0.036 0.1 0.25 0.014 comparison with Schottky cell 306

In/PPEI/PVP(TPD)/ITOb 530 0.038 8 0.3 0.075 λ ) 500 nm 301

Au/CuPc/DM-PTCDI/ITO 415 1.9 100 0.41 0.33 307

Au/ZnPc/DM-PTCDI/ITO 276 0.78 100 0.26 0.07 solvent washed 307380 2.11 0.36 0.29 twice sublimed405 3.53 0.30 0.43 dDIME/dZnPc ) 25/50 nm385 2.17 0.30 0.25 dDIME/dZnPc ) 100/100 nm

Au/H2Pc/PTCBI/ITO 580 0.75 100 0.24 0.10 307

Au/TPP/DM-PTCDI/ITOc 490 0.21 100 0.18 0.019 310Au/TBP/DM-PTCDI/ITO 103 2.57 0.27 0.07Au/H2Hc/DM-PTCDI/ITO 280 0.21 0.37 0.023

Au/H2Pc/DM-PTCDI/ITO 400 2.7 78 0.56 0.76 302(Au/H2Pc/DM-PTCDI)2/ITO 800 0.75 0.21 0.16 tandem with AudAu/(H2Pc/DM-PTCDI)2/ITO 300 0.55 0.35 0.08 tandem without Au

Al/C60/TiOPc/ITO 420 4 × 10-4 10-2 0.20 0.35 illumination through Al 312Al/C60/VOPc/ITO 530 5.4 × 10-5 0.16 0.099 low Pinc at λ ) 720 nm

Al/ C60/PPV/ITOe 800 3.8 × 10-3 0.25 0.48 0.58 PPV spun on 311

a Contacts inferred from text, not explicitly stated. b PPEO: 3,4,9,10-perylenetetracarboxyl-N,N′-bis(phenylethylimide).PVP(TPD): 55 wt % N,N′-diphenyl-N,N′-ditolylbenzidine in poly(vinylpyridine) spin coated onto ITO surface prior to PPEIdeposition. c TPP: 5,10,15,20-21H,31H-tetraphenylporphyrin. TBP: tetrabenzoporphyrin (29H,31H-tetrabenzo[2,3-b:2′,3′-g:2′′,3′′-l:2′′′,3′′′-q]porphyrazine). H2Nc: metal-free naphthalocyanine (29H,31H-tetranaphtho[2,3-b:2′,3′-g:2′′,3′′-l:2′′′,3′′′-q]porphyrazine).d Comparison of single layer cell with tandem cells with and without interstitial 2 nm thick Au layer separating organic bilayers.e PPV: poly(phenylene vinylene).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1869

Page 78: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

temperature was lowered. That is, as observed forthe OMBD growth of PTCDA111 (see section 3.3.3),

the film surface becomes quite smooth as Tsub islowered into the quasi-epitaxial (i.e., kineticallycontrolled) growth regime.The implication of the these results is that contin-

ued improvements in ηP can be obtained for PV cellsgrown by OMBD at Tsub < 300 K under UHVconditions where an extremely high degree of crys-talline order and material purity can be assured.However, it is equally apparent that the resultingmaterials grown in this manner are unlikely toperform at levels which will make them attractivefor use in practical solar energy conversion applica-tions. For this purpose, novel multiple quantum wellstructures such as those discussed in sections 4.2 and4.3 which optimize the overlap between the optical,excitonic and carrier distributions provide the mostlikely means for achieving the long sought-after goalof low cost power generation via the use of efficientthin film molecular organic PV cells.34

A promising, alternative use of organic thin filmsin PVs is as fluorescent down converters layered onthe surface of conventional solar cells. It is wellknown that the efficiency of conventional Si photo-diodes and solar cells decreases rapidly in the near-UV range, while numerous organic thin films exhibitstrong UV absorption followed by radiation at lowerenergies (known as “down conversion”). The downconverted light can be more readily detected by theSi solar cell than the original UV radiation, thus adown-converting organic thin film deposited on thesurface of a conventional solar cell effectively in-creases the solar cell efficiency in the UV range. Thedown conversion is due to the Frank-Condon redshift from absorption to luminescence characteristicof organic molecules. If the film coating thickness isappropriately adjusted, this same film can also beemployed as an antireflection coating for light in thevisible spectral region.317

Experimental and theoretical studies of organicfluorescent down converters for Si solar cells havebeen pursued for almost 20 years.318,319 In the initialstudies, thick (2-5 µm) layers of laser dye moleculesin sol-gel or in plastic matrices were used, andwavelength down conversion (from 400 to 510 nm)with efficiencies exceeding 90% was demonstrated.320Later, deposition of “monocompositional” solid filmsof coronene318 and metachrome321 on CCD sensorsurfaces resulted in the commercial production of UVsensitivity-enhanced cameras. Further studies ofsuccessful utilization of such wavelength down con-verters combined with Si photodiodes have beenlimited to tris(8-hydroxyquinoline)aluminum (Alq3)films.317 Alq3 has a strong absorption at λ < 370 nm(R > 105 cm-1), with a photoluminescence peak at λ) 530 nm (where the Si PV cell efficiency is high)and has a thin-film photoluminescence efficiencyexceeding 30%.322 An Alq3 film deposited directly ontop of a conventional Si solar cell has, therefore, beenobserved to increase the UV collection efficiency ofthe cell by an order of magnitude.317 A similarorganic coating with a 100% conversion efficiencywould increase the solar cell efficiency by as muchas 11% under AM0 illumination. Hence, it is prob-able that new organic thin film materials will leadto the realization of UV-enhanced Si photodiodes and

Figure 5-8. Inverse quantum yield (1/η) at differentapplied fields vs inverse absorption coefficient (1/R). (a) Adiffusion length of LD ) 225 ( 15 nm is found for E ) 2.10to 1.99 eV region. Inset: The data replotted vs 1/η. At otherwavelengths, diffusion lengths observed are (b) LD ) 88 (6 nm for absorption between E ) 2.60 and 2.36 eV and (c)LD ) 79 ( 7 nm for the E )3.2 eV absorption peak weredetermined from linear least squares fits to the data (fromref 172).

1870 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 79: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

solar energy converters with efficiencies sufficientlyhigh to suit many commercial applications.

5.3. Molecular Organic Light-Emitting Devices

Efficient electroluminescence (EL) from an organicsolid was first demonstrated in large (50 µm to 1 mmthick) single crystals of anthracene.323,324 Althoughthese devices showed a high EL quantum efficiencyof up to 8%, they were impractical due to the largeapplied voltage (>1000 V) required to inject electronsand holes into the crystal. More recently, practicalorganic light-emitting devices (OLEDs) consisting ofsequentially vacuum-deposited layers of hole- andelectron-transporting molecular materials31 have beendemonstrated with active device thicknesses of onlya few hundred Ångstroms. In contrast to electronicdevices where crystalline order results in a highcarrier mobility, and hence a good device performance(e.g., in organic TFTs or OI-HJ devices), vacuum-deposited OLEDs typically consist of amorphous thinfilms. The amorphous structure leads to a highradiative recombination efficiency of Frenkel excitonsdue to a reduction in quenching from internal con-version processes linked to highly coupled, crystallineorganic systems. These heterojunction devices haveresulted in a decrease in operating voltage to <10 Vwith a brightness of several hundred candela permeter squared which is adequate for most videodisplay applications.A schematic cross section of a conventional, mo-

lecular OLED with three organic layers (a doubleheterostructure) is shown in Figure 5-9. The top,ohmic, electron-injecting electrode consists of a lowwork function metal alloy, typically Mg-Ag or Li-Al,deposited by vacuum evaporation. The bottom, hole-injecting, electrode is typically a thin film of thetransparent semiconductor indium tin oxide (ITO),deposited onto the substrate by sputtering or electronbeam evaporation. Light is emitted through thiselectrode when the device is operated in “forwardbias”, i.e., with the ITO biased positive with respectto the top electrode. The next layer deposited is ahole transporting layer (HTL) of a material such asN,N′-diphenyl-N,N′-bis(3-methylphenyl)-1,1′-biphenyl-

4,4′-diamine (TPD), shown with other widely usedmolecules in Figure 5-10. This is followed by thelight emitting (EL) layer consisting of, for example,Alq′2OPh for blue light emission, or Alq3 for greenlight. An electron transporting layer (ETL), consist-ing of a material such as TAZ (Figure 5-10) completesthe double heterostructure. Doping a small amount(∼1 mol %) of a fluorescent dye into the EL layer cansignificantly improve the device efficiency, and >5lm/W in the green is now routinely obtained.325Typical deposition rates of all organic molecularlayers is from 1 to 5 Å/s. Examples of molecules usedfor red (R), green (G), and blue (B) light emission arealso shown in Figure 5-10, along with four examplesof charge-transporting molecules. In many cases(Alq3, for example) the functions of electron trans-porter and light emitter may be combined in a singlelayer, resulting in “single heterostructure” devices.There are many variations on the single and double

heterostructures, including hybrid materials combi-nations such as Alq3 deposited on the polymer holeinjecting and luminescent material, poly(p-phenylenevinylene), or PPV.326 In addition, organic/inorganicOLEDs based on the layered perovskite semiconduc-tor, (C6H5C2H4NH3)2PbI4, or PAPI, used as the holeinjection layer in combination with the electroninjecting oxadiazole derivative, OXD7, were shown327to exhibit very narrow spectral emission centered atλ ) 520 nm.While OLEDs based on vacuum-deposited molec-

ular materials31,32,264,325,328-330 or on polymers331-333

have both been investigated,330 the discussion insections 2 and 3 clearly shows that vacuum deposi-tion offers the advantages of better control over layerthickness and uniformity, and access to reproducible,higher purity starting materialssall critical param-eters for displays. In addition, since all of the devicelayers are vacuum deposited at room temperature,OLED fabrication is compatible with conventionalsemiconductor processes. Failure mechanisms ofsuch devices are primarily related to dark spotformation at the contact/organic thin film interfaces.These effects have been extensively studied,334-337

and recently continuous operating lifetimes of vacuum-deposited green OLEDs at video brightness havebeen shown to exceed 20 000 h,338,339 making themsufficiently long-lived for commercial flat panel dis-play applications.The devices have also been found to be vulnerable

to structural instabilities as the temperature isincreased beyond the glass transition temperature(Tg) of one of the layers composing the OLEDheterostructure.340-342 The OLED structure undertemperature stress was thoroughly studied usingglancing incidence X-ray,341 employing the sametechniques and tools discussed in section 3.3.2.3 aswell as via energy dispersive X-ray diffraction,342 withboth investigations resulting in essentially similarresults. Here, we discuss briefly the results of Fenteret al.,341 since they illustrate some of the advantagesto using vacuum-deposited materials in achievingprecise structural control, and also provide informa-tion as to the degradation under thermal stress ofOLED heterostructures.

Figure 5-9. Conventional organic light emitting device(OLED) double heterostructure showing the contacts, theelectron transport layer (ETL), light-emitting layer (EL),and hole transport layer (HTL). The total organic thin filmthickness is typically ∼1000 Å.

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1871

Page 80: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

In Figure 5-11a, the X-ray specular reflectivity ofa series of thin film structures (thin and bold linesare experimental data and theoretical fits, respec-tively), is shown. These structures are (i) TPD on aSi substrate, (ii) Alq3 on Si, and (iii) Alq3/TPD on Si.The reflectivity is dominated by the Fresnel reflec-tivity of a highly flat interface, whose intensityasymptotically follows 1/Q4, where Q is the X-raymomentum transfer. A least-squares fit to the data(bold lines, Figure 5-11a) using standard Fourieranalysis techniques,343 provides the electron densityprofiles of the layers (Figure 5-11b). These fitssuggest that the mean-square interface roughness inall three structures is <4.5 Å. This roughness iscomparable to the dimensions of a single moleculeand is only ∼1% of the total film thickness, providingevidence that vacuum deposition of organic films canresult in extraordinarily flat (on the molecular scale)

films and interfaces necessary to realize many, highperformance optoelectronic devices.Fourier transforms (FT) of the reflectivity oscilla-

tions provide extremely clean data (Figure 5-11c) asa result of the flatness of all interfaces. The layerthicknesses obtained from these transforms for boththe single-layer structures and the compound struc-ture are clearly indicated by the peaks in the figureand correspond very well to the thicknesses mea-sured during growth using a quartz crystal microbal-ance. Using such precision data, Fenter et al.341 wereable to accurately measure such film properties asthermal expansion coefficients and degradation ofOLED structures under thermal stress. An exampleof the changes which occur to a TPD film as itstemperature is elevated above Tg ≈ 60 °C is shownin Figure 5-12. The reflectivity spectra in Figure5-12a for a TPD/Si film clearly indicate an attenua-

Figure 5-10. Structural formulae for several compounds typically used in OLEDs.

Figure 5-11. (a) X-ray specular reflectivity (thin lines) of (i) Alq3, (ii) TPD and (iii) Alq3/TPD films grown on Si substrates.Corresponding fits to the data are shown by the bold lines, with data and fits offset for clarity. (b) Electron density profilesderived for each film structure. (c) Fourier transform (FT) of the data shown in a (from ref 341).

1872 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 81: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

tion of the signal as temperature is increased, withirreversible changes occurring at T > 85 °C. Bytaking the FT of these data, the film thickness canbe directly tracked as the film is thermally cycled.This is indicated in Figure 5-12, parts b and c, wherethe film is first heated to 85 °C, then cooled, and thenreheated to 105 °C. It is observed that in the firstcycle, the film expands by 9 Å, from which a thermalexpansion coefficient of R ) 1.2 × 10-4/K is inferred.Upon cooling, it is found that the film thicknessactually increases slightly, along with a slight de-crease in FT peak intensity. During the secondcycling above the melting point of TPD, the FTintensity decreases by a factor of 102, although thefilm thickness decreases only slightly. The FT in-tensity decrease rules out the possibility that the filmhas simply densified due to the thermal cycle, sug-gesting instead that laterally inhomogeneous des-orption occurs, leaving small TPD islands whosethickness is nearly unchanged from that of theoriginal film. Indeed, the X-ray data do not supportthe presence of any recrystallization,340 but ratheronly that morphological changes are occurring.Further X-ray studies of the compound Alq3/TPD/

Si structure along these same lines were used todevelop a complete picture of OLED structural failureduring thermal stress.341 It was concluded that theprimary driving force for the degradation of Alq3/TPDheterostructures, and which is undoubtedly similarlyactive in many other organic heterojunction systems,is the large thermal expansion of one of the compo-nents (TPD) associated with its glass transition,creating catastrophic strain release at the heteroint-erface between the materials. This conclusion issupported by the observation that Alq3/TPD struc-tures grown at elevated temperatures (Tsub ) 73 °C)exhibited better morphology and greater thermalstability than those grown at room temperature.334

Thermally induced failure of OLEDs is reducedunder normal device operating conditions by increas-ing the Tg of all materials incorporated into thestructure. For this reason, TPD has now largely beenreplaced by R-NPD (Tg ≈ 110 °C) in vacuum depos-ited OLEDs (see Figure 5-10). Tokito et al.344 haveshown that a tetramer of triphenylamine (TPTE)with Tg ) 130 °C can also be used as an efficient hole-transporting material. Such TPTE/Alq3-based de-vices were demonstrated to have stable operation atT e Tg. Indeed, storage at T ) 150 °C was not foundto result in significant degradation of the devicecharacteristics when operated after returning to roomtemperature. Many other aromatic amines have alsobeen investigated for their suitability as high stabil-ity, high melting temperature hole transport materi-als for OLED applications.295 Finally, we note thatin addition to thermal instabilities and dark spotformation, photochemical instabilities334,345 of theorganic materials can also influence the lifetime ofthese interesting devices.In the remainder of this section, we discuss the

fundamental processes leading to current conductionand electroluminescence in vacuum-deposited OLE-Ds. Discussion of these processes leads to a morecomplete understanding of the relationship betweenfilm structure and device properties. This is followedby a discussion of some of the interesting OLEDarchitectures which are apparently unique to thisnew generation of organic devices.

5.3.1. Current Transport and ElectroluminescenceMechanisms in OLEDs

It has recently been shown32,258,346 that the electri-cal and optical characteristics of vacuum-depositedmolecular OLEDs can be modeled by assuming thatthe fundamental mechanism controlling both currentconduction and EL yield is space charge effects in thepresence of an intrinsic trap distribution. The depthand distribution of the traps is consistent with theirbeing formed by molecular relaxation on ionizationby injected charge carriers. That is, once a chargecarrier (e.g., an electron) localizes on a molecule (i.e.,it is trapped), a charge of opposite polarity (a hole)injected from the counter electrode becomes similarlylocalized on that same molecule. This electron-holepair, whose energy is less than the ionized molecularenergy by their Coulomb binding energy, eventuallyrecombines, emitting either light or heat. Most ofthe recombination occurs within only a few hundredÅngstroms of the HTL/ETL interface determined bythe diffusion length of carriers in the emitting mate-rial. This region is known as the “recombinationzone”.It has been found32,347 that exciton diffusion lengths

in such amorphous thin films as Alq3 are approxi-mately 100-300 Å. Hence, in optimal devices, theAlq3 layer thickness must be only slightly larger thanthis thickness (see below), suggesting that significantcontrol over film deposition as well as on substrateflatness afforded by vacuum deposition of molecularorganics is necessary.A very clear demonstration of the extent of the

recombination zone was provided by Adachi et al.348using a single heterostructure OLED shown sche-

Figure 5-12. (a) Normalized X-ray specular reflectivityfor TPD/Si vs temperature. (b) Film thickness and (c)calculated FT intensity vs temperature. The arrows in band c indicate the thermal history of the sample (from ref341).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1873

Page 82: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

matically in Figure 5-13. For this device, the 500 Åthick hole transport and light-emitting layer con-sisted naphthostyrylamine (NSD), and the 500 Åthick ETL consisted of an oxidiazole derivative.Structural formulae for molecules used in this studyare also shown in Figure 5-13. To determine theextent of the recombination zone, a highly emissivedopant chromophore, a thiadiazole derivative, wasgrown in very thin (10 Å) slabs located at differentdistances from the HTL/ETL interface in differentdevices. The emission spectrum of each of thesedevices was then measured, with the results shownin the figure. For dopant slabs located within ∼100Å of the heterointerface, the emission peaks at 590nm (spectra labeled A and B), characteristic of thespectrum from the thiadiazole derivative. However,when the dopant is moved farther from the interface(spectra C-F), the emission peak shifts to 500 nmdue to recombination in the undoped NSD. Thisresult clearly indicates that virtually all radiativerecombination occurs in the HTL, within 100 Å of theHTL/ETL interface for this heterostructure. Theseconclusions about the width and position of therecombination zone has been confirmed by measure-

ments of exciton dynamics in Alq3 and at Alq3/TPDinterfaces.349,350 Undoubtedly, these conclusions canbe generalized to understand almost all small mo-lecular weight, vacuum-deposited OLEDs.Typically, the power-law dependence of current-on-

voltage observed for OLEDs and expected from thetrapped charge model fits experimental data overmany decades of current and over a wide range oftemperature. However, effects due to low efficiencyof charge injection from the electrodes, particularlythe low work function cathode, can complicate theprocess by adding extra resistance which results ina voltage offset.351In the trapped-charge model, we assume that the

current density is bipolar, in that it consists primarilyof holes injected from the ITO contact into the HTL(e.g., TPD), with electrons injected from the low workfunction cathode (e.g., Mg-Ag) into the ETL (e.g.,Alq3). Of the two carriers, electrons have the lowermobility and hence limit the current conductionprocess. Radiative recombination occurs when holesare injected from the TPD into the Alq3, where,within a short distance (a few hundred Ångstroms)of the TPD/Alq3 heterojunction, they bind to anelectron on an Alq3 molecule forming a short lifetime,Frenkel exciton. Radiative recombination competeswith nonradiative (dissipative) recombination occur-ring at the TPD/Alq3 interface as well as at thecontacts.The power-law dependence of current-on-voltage

(Figure 5-14), commonly observed in both molecularorganic32,258,326,346,351,352 and polymer-based353-355 OLE-Ds, is interpreted in terms of the injection of spacecharge into a trap-filled insulator (Alq3). At very lowapplied voltage, ohmic conductivity dominates, wherethe current density, J, is described by J ) qµnnoV/d.Here, no is the background density of free electronsin Alq3, µn is the electron mobility, and d is the layerthickness. Previous observations of ohmic conductionat extremely low injection currents32 (<0.1 µA/cm2)sets an upper limit to noµn in this expression of 107

Figure 5-13. Top: Chemical structural formulae for theorganic molecules used to determine the width of therecombination region. Center: Schematic of the dopedsingle-heterostructure OLED. Bottom: Electrolumines-cence spectra of six doped single heterostructure OLEDs,with the positions (A-F) of the dopant layers indicated inthe diagram of the OLED (from ref 348).

Figure 5-14. Current as a function of applied voltage ina 270 Å thick TPD, and 550 Å thick Alq3 OLED at varioustemperatures (open symbols). The solid lines show a fit tothe space-charge limited model discussed in the text, usingNt ) 3.1 × 1018 cm-3, and µNLUMO ) 4.8 × 1014 (cm Vs-1). Inset: Temperature dependence of the power lawparameter,m, (I ≈ Vm + 1) giving Tt ) 1780 ( 50 K (fromref 258).

1874 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 83: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(V cm s)-1. At higher injection currents, the fillingof traps below the electron quasi-Fermi level resultsin the current being governed by the density andenergy distribution of the traps. An analyticalexpression relating current to voltage in this trapcharge limited (TCL) regime is obtained by assuming,for simplicity, a continuous exponential energy dis-tribution of traps below the LUMO following356

Here, Nt is the total trap density, ELUMO is the Alq3LUMO energy, and Tt ) Et/k, where Et is thecharacteristic trap energy. The current density isthen given by

where ε is the permittivity and NLUMO is the LUMOdensity of states. It is clear from eq 5.3 that m )Tt/T determines the temperature dependence of theI-V characteristics. A weaker effect is the temper-ature dependence of the mobility, which in organicmaterials empirically follows µn(T) ≈ T-n, where nvaries from 0 to 3. Conduction and scattering byphonons in narrow bands characteristic of organicmolecular solids has been shown to yield a temper-ature dependence357 of µn ≈ T-2. However, smallchanges in the exponent or functional form or eventhe small expected dependence on electric field,chosen for µ(T) do not significantly affect the fit,which is primarily governed by the strong power-lawdependence of eq 5.3. Finally, the temperaturedependence of NLUMO has also been shown to benegligible.258The forward-biased I-V characteristics of TPD/

Alq3 OLEDs at temperatures ranging from 120 to 300K at 20 K intervals have been fit to the modeldescribed above,258,346 as shown by the solid lines inFigure 5-14. For these fits, it was assumed that theOLED current is limited by the 550 Å thick Alq3layer, and hence the contacts play a less significantrole in limiting current transport in optimum deviceswith thin organic layers. With Nt and µnNLUMO asparameters, eq 5.3 fits the data over six decades ofcurrent, with m varying from 6 ( 1 at 300 K to 15 (1 at 120 K. The inset to Figure 5-14 shows thevariation of m with 1/T, which is fit by a straight linecorresponding to Et ) 0.15 ( 0.02 eV. A further testof the validity of the trap-limited model is theprediction from eq 5.3 that V is approximatelyproportional to d2 at constant current, which has alsobeen confirmed for TPD/Alq3 devices.258,346Using the measured value of Tt, the density of traps

can be determined by fitting the temperature de-pendent data to eq 5.3. From this, it was foundthat258,346 Nt ) 3.1 × 1018 cm-3, and µnNLUMO ) 4.8 ×1014 (cm V s)-1. Using time-of-flight measurements,Hoskawa et al.358 determined that µn(300 K) ) (5 (2) × 10-5 cm2/(V s), from which we infer NLUMO ) (1.0( 0.5) × 1019 cm-3. Finally, this model yields no <1011 cm-3, consistent with the insulating nature ofAlq3.

The fact that NLUMO and Nt are roughly equal (andwithin 2 orders of magnitude of the density ofmolecular sites) suggests that each conduction elec-tron is self-trapped on an Alq3 molecule32,258,346 dueto the resulting generation of a polaron, i.e., eachmolecular site serves both as an available electronstate and a trapping site. This is consistent withmolecular orbital calculations258 which indicate thatthe Alq3 anion LUMO is∼0.2 eV lower in energy thanthe neutral statesa value approximately equal to theneutral polaron energy, Et. Hence, an electron andhole which localize on the same quinolate ligand ofan Alq3 molecule (which are the sites of the calculatedHOMO and LUMO as shown in Figure 5-15), arelowered in energy from their conduction states by theCoulomb binding energy of the resulting Frenkelexciton. The exponential distribution of energies, assuggested by the above analysis, is due to thepresence of a high density of intramolecular phononsat room temperature leading to a distribution ofnondegenerate molecular conformations. Evidencefor the effects of phonons is also apparent from thebroad EL and PL emission spectra32,347 which are atsignificantly lower energies from the absorptionspectrum of Alq3 due to the large Franck-Condonshift. In section 5.3.2, we show that this substantialred shift in luminescence is useful in transparentOLEDs which have a wide range of device applica-tions.256

These relaxation processes can be put in thecontext of the HOMO-LUMO energy diagram for anOLED shown in Figure 5-16. Here, electrons areinjected from the cathode into the Alq3, and holes are

Nt(E) ) (Nt/kTt) exp[(E - ELUMO)/kTt] (5.2)

JTCL ) NLUMOµnq(1-m)[ εm

Nt(m + 1)]m ×(2m + 1m + 1 )(m+1) V(m+1)

d(2m+1)(5.3)

Figure 5-15. (Bottom) The location of the filled orbitals(HOMO states) in a molecule of Alq3 and (top) the locationof the empty orbitals (LUMO states) (from ref 258).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1875

Page 84: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

injected from the TPD from the opposite electrode.These carriers hop from molecule to molecule, withthe electrons arriving in the vicinity of the ETL/HTLinterface before localizing on an Alq3 molecule. Thislocalization is a “self-trapping”, or polaron generationprocess which lowers the energy of the molecule inits anionic form to well below ELUMO. Subsequently,a carrier of opposite sign (in this example, a hole)localizes on the anion, further decreasing the now-excited molecular state by the Coulomb pair bindingenergy. At some time later equal to the excitonlifetime, the pair recombines (either radiatively,generating light, or nonradiatively, generating heat)and the molecule is free to accept the next injectedcarrier.The role of self-trapping has in fact been extended

to quantitatively explain the mechanisms for EL.258,346In Figure 5-17, the EL quantum efficiency at aconstant current of 50 µA (solid symbols) is plottedas a function of temperature. Now, the EL flux (ΦEL)is proportional to the rate at which electrons in thefilled trap states recombine with holes in the Alq3HOMO (at energy EHOMO) via an intermediate, re-

laxed Frenkel exciton state. Since holes (of densityp(x)) are minority carriers in Alq3, the recombinationrate per volume as a function of the distance fromthe electron injecting Mg:Ag electrode at x is

where τp(x) is the hole lifetime in the Alq3 layer,which can be determined from steady state recom-bination statistics.359 Also, p(x) is found using

where Dp ) µpkT/q is the hole diffusion constant, µpis the hole mobility and F(x) is the electric field. Thetotal recombination rate is R = p(d)µpF(d). Since theratio of radiative to total recombination is propor-tional to the photoluminescence (PL) efficiency, ηPL,in Alq3, the temperature dependent EL flux is

The factor, R, includes all losses that do not exhibita significant temperature dependence (e.g., due tospin statistics, small-energy nonradiative transitions,etc.). Here, ηPL(T) is shown in the inset in Figure5-17. The forward-scattered external PL quantumefficiency, which can be directly compared to the ELin an OLED, was measured to be 7%.Assuming µp ) 0.01µn, Ep - EHOMO ) 20 ( 1 meV

at x ) d, R ) 0.25 and NHOMO ) 0.07NLUMO, the ELdata of Figure 5-17 have been fit (solid line) to eq5.6. The EL intensity increases with decreasingtemperature due to a decrease in the hole lifetime-diffusion length product. The increase saturates atthe lowest temperatures due to the reduced holeinjection into the Alq3 as a result of a decrease inp(d). The calculated room temperature external ELquantum efficiency (η ) qΦEL/J) obtained from thisfit is ∼1%, consistent with the measured value of0.5-1%.This study clearly suggests a link between the filled

traps that give rise to the I-V characteristics andthe EL intensity, providing detailed insight into thephysical origin of EL in OLEDs. One consequencedrawn of this analysis is that only ∼1% of the Alq3molecules are optically active under even the highestdrive currents. This is not surprising in that veryhigh currents should lead to a large electron spacecharge in the film which would serve to halt (viaCoulombic interactions) further current injection.This analysis is essentially a “one-electron” picturewhich does not account for electron-electron interac-tions. Nevertheless, the model suggests that optimalperformance can be achieved by lightly doping aconducting host molecular film with ∼1% of a differ-ent, highly luminescent guest molecule to achievemaximum EL efficiency. This effect was observed264for several different molecular dopants in Alq3,including360,361 the red-emitting tetraphenylporphine(TPP). It was found that, as predicted, maximum redluminescence efficiency (with a peak wavelength ofλ ) 655 nm, Figure 5-18) is achieved for TPP dopingsof ∼1% (by mol wt), with the device red-emissionefficiency decreasing at both higher and lower con-

Figure 5-16. Schematic energy level diagram showing theprocess of charge transport and subsequent relaxation oftrapped electrons in Alq3 with a minority hole, to form adistribution of Frenkel excitons in the Alq3 bandgap. Theseself-trapped excitons subsequently recombine to generateelectroluminescence. Electrons are indicated by shadedcircles in the LUMO band, and holes by open circles in theHOMO band. Fp and Fn indicate the quasi-Fermi levels forholes and electrons, respectively (from ref 258).

Figure 5-17. Dependence of electroluminescence efficiencyon temperature of a 270 Å thick TPD and 550 Å thick Alq3OLED (closed squares), and fit to a model assuming thattrapped charge gives rise to EL (line). Inset: Dependenceof photoluminescence efficiency on temperature of a 1000Å thick Alq3 film vacuum evaporated on a UV-transparentquartz substrate (from ref 258).

r(x) ) p(x)/τp(x) (5.4)

p(x)τp(x)

) Dpd2p(x)

dx2- µp

ddx[p(x)F(x)] (5.5)

ΦEL(T) ) RηPL(T)R ) RηPL(T)p(d)µpF(d) (5.6)

1876 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 85: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

centrations of TPP in Alq3. The energy transfer fromAlq3 into TPP is complete at this level of doping, asshown by the lack of measurable Alq3 emission(typically observed with a peak wavelength at ∼530nm), falling off as the TPP concentration is furtherreduced. In Figure 5-18 is shown the OLED spec-trum for very lightly doped TPP (<0.5%) in Alq3,where incomplete energy transfer is evident from theemergence of Alq3 emission (peak B) at high injectionlevels.362 For concentrations >1%, only the two TPPemission features (peaks C and D) are evident at allinjection levels.360Other models have also been suggested to explain

the current conduction in OLEDs. For example, theI-V characteristics in OLEDs have been fit over twoto three decades of current by a Schottky-type ther-mionic emission model,363,364 although a large dis-crepancy (2-6 orders of magnitude) between themeasured and predicted values of the saturationcurrent was observed. A model which assumes thatcurrent is limited by tunneling from the contacts intothe organic layer258,365 can also be made to fit theobserved data for polymeric OLEDs over one to twodecades of current in the high-field regime.The efficiency of OLEDsmay be strongly influenced

by subtle changes in the organic film structure. Forexample, Joswick et al. systematically investigatedthe effect of film crystallinity on EL device perfor-mance.366 Their results indicate that increasing thepolycrystallinity of the EL layer made from conju-gated oligomers dramatically decreases both carriermobility and EL quantum efficiency. Such effects arealmost certainly also present in molecular organic

films. Amorphous, glassy films, or alternativelyhighly crystalline films such as those grown byOMBD, are therefore desirable for achieving efficient,low-voltage OLEDs.Finally, the trap-filled conduction model is clearly

valid only when the cathode and anode contacts areohmic, with a sufficiently low resistance such thatthe current is limited primarily by transport throughthe thin films.258 This condition is met in optimalOLEDs where the Alq3 layer thickness is ∼500-700Å, and is approximately double that of the HTLthickness.367 Furthermore, the formation of darkspot defects at the low work function metal/ETLinterface ultimately determines the operational life-time of the devices.336,337,345 Consequently, under-standing the fundamental properties of the organic/metal interface368 and other organic thin filmsystems369,370 is of the greatest importance in OLEDs,TFTs, and virtually all electrically active organicdevices, and hence has been of considerable recentinterest. While it is generally thought that low workfunction metals are required to obtain a low electroninjecting barrier at the ETL/cathode interface, it isapparent that mechanisms resulting in ohmic contactbehavior are quite subtle, involving significant metal/organic interdiffusion,371 and chemical reactions be-tween the species.372 These processes can generatea high density of interface defect states which domi-nate the injection properties of the contact into theorganic film. In photoelectron spectrographic studiesof metal/PTCDA contacts, Hirose et al.369 have foundthat these defect states may be responsible for thelow resistance contacts observed, as opposed tosimple work function and energy band alignmentarguments.The optimization of contacts in current injection

devices such as OLEDs, organic lasers, and transis-tors is possibly one of the most important consider-ations in making practical organic-based electronicdevices, and hence continued studies of the metal/organic interface are of the utmost importance.

5.3.2. Organic Light-Emitting Device Structures andApplications

One of the principal reasons that OLED technologyhas attracted such interest in recent years has beenthe prospect for using these devices in full color, red-green-blue (RGB) emissive displays which mighteventually replace active matrix liquid crystal dis-plays. Several techniques have been proposed forgenerating RGB displays,373 as shown in Figure 5-19.Simple, side-by-side photolithographic patterningcombined with shadow masking allowed Wu et al.374to fabricate red, green, and blue polymer OLEDs onthe same substrate, although at very low resolution.This approach, shown in Figure 5-19a, while archi-tecturally obvious, can result in a high-cost processwhere growth and fabrication of the three colorsegments must be alternated without damage to thepredeposited and prefabricated subpixels. Alterna-tively, optical filtering of white OLEDs can produceacceptable red, green, and blue emission (Figure5-19b), but at a great cost to device efficiency due tothe large amount of light absorbed in the filters.Several demonstrations of bright, white-light emit-

Figure 5-18. Output from a Mg-Ag/Alq3:TPP/TPD/ITOorganic light-emitting device as a function of drive current.For this device, the TPP concentration in Alq3 is very small(∼0.5% by wt) to show effects of energy transfer saturation.The drive currents for the 0.8 mm2 device are (starting fromthe highest intensity curve) 3 mA, with the drive currentfor each successive curve reduced by approximately a factorof 2 from the adjacent, higher intensity spectrum. PeaksC and D are characteristic of TPP, whereas peak B ischaracteristic of Alq3. It is observed that the intensity ofpeak B scales with pump intensity, indicating a saturationof luminescence from the limited concentration of TPPmolecules. Furthermore, the ratio of intensities of peak Cto B also decreases as drive current is increased above 500µA providing further evidence for saturation of energytransfer. For TPP concentrations exceeding 1%, energytransfer is ∼100%, completely eliminating emission fromAlq3 at all practical OLED output powers (from ref 362).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1877

Page 86: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

ting OLEDs375,376 utilized multiple EL layers wherethe molecular composition was varied to give thespectral content necessary for white emission. At-tempts at using rare-earth complexes to build OLEDswhich emit “white” light concentrated into the red,green, and blue spectral bands, however, greatlyreduced the efficiency of the OLED itself.377Less efficiency is wasted by using a single blue

OLED378 to pump organic wavelength “down convert-ers” (Figure 5-19c). Each down converter consists ofa prepatterned film of fluorescent material whichefficiently absorbs blue light and reemits the energyas either green or red light, depending on thecompound used.379,380 Luminescent organic systemscan have a conversion quantum efficiency approach-ing 100%, although the power efficiency is reducedsince the energy of the emitted photon is less thanthat of the absorbed photon. This is particularlyimportant in the generation of red light by a bluepump, which results in a maximum down-conversionpower efficiency of ∼50%, assuming the down-conversion quantum efficiency is 100%. Neverthe-less, a low-resolution display using down convertershas recently been reported.381Use of OLEDs which are intrinsically color tunable

obviates the need for separate RGB elements, result-

ing in a 3-fold improvement in resolution and displayfill-factor as compared to the various side-by-sidearchitectures. Only one OLED structure is grownover the entire area of the display, although thissimplicity is somewhat offset by additional complex-ity required in the drive circuit, which must simul-taneously control color, brightness, and gray scale.For example, polarity-dependent, two-color operationhas been obtained from OLEDs doped in differentregions of the heterostructure with molecular specieswhich emit at different wavelengths.382 In this case,the color is tuned by varying the polarity and themagnitude of the applied voltage; higher voltageresults in more emission from the blue emittingcomponent, while also resulting in higher overallbrightness due to increased current injection into thedevice. To reduce this problem, emission intensitycan be controlled by using a pulsed current sourceat reduced duty cycles. In a color display, therefore,high drive voltages at very low duty cycles may benecessary for blue pixels, which is likely to enhancedegradation. Note that molecular organic color-blendstructures have not yet been demonstrated which cancontinuously access all regions of the color spectrum.In a high-resolution display, color tuning the pixelswhile independently controlling brightness and greyscale appears to be problematic.Organic thin films, however, allow for the realiza-

tion of completely new display concepts which, amongother advantages, may also lead to the practicalrealization of low cost RGB displays. For example,Gu et al.256,383 demonstrated transparent OLEDs (orTOLEDs), useful for transparent head-up, or veryhigh contrast displays (inset: Figure 5-20). Thesedevices employ the large Franck-Condon shift char-acteristic of many organic molecules196,197,347 whichresult in significant red shifts of the emission relativeto the absorption spectrum. Hence, these films aretransparent to their own emitted light. In Figure5-20, we show the spectral output from a blue doubleheterostructure TOLED373,384 consisting of a 500 Åthick TPD layer used as the HTL, followed by a 200Å thick Alq2OPh as the emission layer, a 500 Å thickAlq3 ETL, and a cathode contact consisting of 100 ÅMg-Ag followed by a sputter-coated ITO contact. Thelight emission from the top surface of the TOLED

Figure 5-19. Schemes for creating full color displays basedon organic light-emitting devices placed in a side-by-sideconfiguration. The methods are (a) closely positioned,separate red, green, and blue emissive pixels; (b) a singlewhite OLED positioned above red, green, and blue colorfilters; (c) a single blue or UV OLED positioned above red,green, and blue fluorescent down converting layers. In thecase of the red pixel, it may be necessary to first generategreen through fluorescent down conversion, followed by asecond step to red; (d) a single, very broad spectral emissionOLED positioned above dielectric mirror stacks to providemicrocavities tuned with their optical band pass in the red,green, and blue spectral regions. (Figure courtesy of Y.Sato.)

Figure 5-20. Spectral emission from a blue, transparentorganic light emitting device (TOLED) from the top andsubstrate surfaces. Inset: Structure of the double hetero-structure TOLED (from ref 384).

1878 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 87: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(with a total efficiency of 3.65%) is somewhat blueshifted with respect to the substrate emission due toabsorption in the Mg-Ag contact.In addition to being a transparent light emitter,

the top ITO surface of the TOLED can serve as thehole injecting electrode for a second TOLED built ontop of the first device. Each device in the stack isthen independently addressable and can be tailoredto emit its own color through the adjacent transpar-ent organic layers, the transparent contacts and theglass substrate. This allows the entire area of thevertically stacked pixel to emit any mixture of thetwo primary TOLED colors. Recently, both two- andthree-color SOLEDs have been demonstrated,361,385with the three-color device shown schematically inFigure 5-21a. Since each color element in the stackis independently addressable, a stacked OLED(SOLED) display can be built with independentcontrol of brightness, color, and gray scale. In theSOLED in Figure 5-21a, the blue double heterostruc-ture (with a structure similar to that shown in Figure5-20 except with the TPD replaced with the improvedstability, R-NPD) was grown at the bottom, and isseparated by the second (green) subpixel by a com-mon, semitransparent Mg-Ag electrode. The conven-tional green emitter was then grown with its Alq3ETL and emission layer sharing this same cathode,followed by a R-NPD HTL. Since this structurerequires the somewhat destructive sputter depositionof ITO on top of the HTL to form the anode contact,a thin layer of PTCDA is first deposited onto theR-NPD prior to ITO deposition. PTCDA serves twopurposes: (i) Since it is an extremely robust molecule,it prevents sputter damage to the underlying R-NPD,and (ii) its high hole mobility37,172 improves carrierinjection into the green emitter. Note that thecontact properties for organics deposited onto thesurface of the Mg-Ag are considerably different thanfrom those where the Mg-Ag is deposited onto theorganic thin film. Indeed, in the later case, thecontact appears to have significantly lower resistancethan in the former, presumably due to chemistrywhich occurs during the condensation of highlyenergetic atoms onto the surface of the organicmaterial.367

To complete the three-color device, a third R-NPDlayer was deposited onto the sputtered ITO contact,followed by a 3% TPP:Alq3 ETL and light emittinglayer. The final cathode was a conventional, non-transparent Mg-Ag layer.The electroluminescence spectrum from the several

different emitters in the stack is shown in Figure5-21b. The quantum efficiencies for the red, green,and blue subpixel elements were 0.3%, 0.55%, and1.6%, with output luminances of 35, 70, and >200cd/m2, respectively, adequate for many display ap-plications. Note that these spectra are significantlydifferent than those obtained for three discreteOLEDs using the same emitting molecules.385 Thisis due to the existence of optical cavities formed bythe several different organic dielectric layers and thereflective, semitransparent metal contacts comprisingthe stack. These so-called “microcavity effects” havebeen extensively studied386-393 and have been foundto not only affect the spectral content of the emitted

light, but also its angular distribution. While sucheffects are often undesirable, it has recently beenproposed388 that multicolor displays consisting ofOLEDs tuned by external dielectric reflecting layerscan be one approach to achieving full color, as shownin Figure 5-19d.As discussed extensively in this review, organic

thin films have high mechanical flexibility, and donot require lattice matching with the substrate. Thisproperty creates new possibilities for the deviceengineer, such as the 12-layer SOLED shown inFigure 5-21, which is a unique combination of amor-phous and crystalline organic films, inorganic con-ductive oxides and metal films all comprising asingle, integrated and efficient structure. Currently,

Figure 5-21. (a) Exploded schematic view of a three-colorstacked organic light emitting device (SOLED) showing thevarious material layers and the biasing scheme. The totalthickness of the contacts and organic layers is ∼0.8 µm.(b) The output spectra from the three separate subpixelsin the SOLED. Under operation, the total spectrum is alinear superposition of the separate spectra of the indi-vidual subpixels (from ref 385).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1879

Page 88: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

it is not possible to fabricate such heterogeneouslayered structures based on inorganic semiconductormaterial systems.In other applications of organic-based light emit-

ters, polarized EL, with potential applications foroptical memories, has been obtained by either orient-ing the organic molecules on the substrate sur-face394,395 or confining the EL layer in a multilayerstructure.148,261 In addition, flexible, molecular OLE-Ds as shown in Figure 5-22, have been demonstratedon ITO-coated polyester substrates,256,396 whichcomplement earlier demonstrations of flexible poly-meric OLEDs,331,397 creating the potential for roll-upor conformable displays on curved surfaces. Vacuumdeposition on plastic sheets also allows mass-produc-tion of large-area OLEDs via roll-to-roll processing.The successful development of such a process couldresult in extremely low cost manufacture.We conclude this section by noting that work

published by over 80 laboratories worldwide398-400 inthe last few years has shown that arrays of vacuum-deposited, green OLEDs are sufficiently bright andlong-lived to be a realistic alternative to conventionaltechnologies such as back-lit liquid crystal displays,and indeed are currently being adopted for earlyentry into commercial markets.30,401 OLEDs basedon vacuum-deposited organic thin films exhibit uniqueproperties of transparency, flexibility, and simplicityof manufacture which are unobtainable using con-ventional, inorganic semiconductors. These proper-ties enable new display applications. The intenseworldwide effort to exploit OLED technology appearspoised to succeed in producing an entirely newgeneration of low-cost flat panel displays with novelproperties inaccessible with existing display concepts.

5.4. Thin-Film Transistors

One area where remarkable progress has beenmade during the last two years265-270,402-404 has been

in the understanding and fabrication of organic thinfilm transistors (TFTs). Interest in these devicesstems from their potential use in very low costdisplay back plane active transistor matrices, and inlow-cost (albeit low performance) electronics applica-tions such as in credit card ID encoding, etc.405Organic TFTs have the potential for roll-to-rollproduction due to their compatibility with low-temperature fabrication processes. Indeed, if workin this area is successful, organic TFTs may one dayreplace such devices currently based on amorphousSi which require elevated temperature deposition(typically ∼200-500 °C), making them incompatiblewith integration on ultralow cost polymer substrates.Charge carrier mobility, the characteristic of great-

est importance to TFT performance, is determinedby the degree of π-orbital overlap between adjacentmolecules forming the conducting channel from tran-sistor source to drain. In addition, to minimize thedark conductivity of the transistor channel, thebackground free carrier concentration must be mini-mized, placing stringent requirements on materialpurity (where defects and impurities give rise toshallow donor or acceptor states). For these reasons,it has been suggested268 that the most promisingapproach to achieving the required high mobility andlow impurity concentration is via vacuum depositionof highly pure planar aromatic compounds such aspentacene,269 the oligothiophenes (e.g., R-hexathie-nylene, R-6T),266,270 perylene and naphthalene-basedcompounds,224,406 and the Pc’s.407

The structure of a typical organic TFT is shownschematically in Figure 5-23a. Here, an organic thinfilm which serves as the transistor channel is depos-ited to a thickness of 1000-2000 Å onto an oxidized,conductive Si surface between ohmic metal (e.g., Au)or polymer (e.g., polyaniline) contact fingers. Thecontact fingers serve as the source and drain elec-trodes of width W, and separation (i.e., channellength) L. The Si substrate provides the gate contact.Conduction occurs at the organic/SiO2 interface whena bias, VGS, whose polarity depends on the n or p-typecharacter of the organic thin film, is placed betweenthe gate and grounded source contact. For p-channelaccumulation mode TFTs, for example, applicationof a sufficiently high VGS lowers the surface Fermi-energy of the film below the HOMO, thus creating aconducting pathway between source and drain, withthe total current determined by the drain voltage,-VDS. Inversion mode and n-channel devices workin much the same fashion, although the sign of thegate and drain potentials are reversed.While the best device performance to date269 has

been achieved using the structure in Figure 5-23a,we note that all of the gate contacts in a multitran-sistor integrated circuit using such a device geometrywill be shorted. For practical application, the sub-strate must be an insulator such as a polymer sheetor glass. In this case, a conducting gate electrode isfirst deposited and patterned, followed by depositionof the gate insulator, which can be either a polymer,a vacuum-deposited organic film, or an inorganicmaterial such as SiO2 or SiNx. After the gateinsulator is deposited, further processing follows the

Figure 5-22. Photograph of an array of nine, unpackaged1 cm2 vacuum-deposited, nonpolymeric flexible OLEDs.One device (near top of photo) contacted with the probearm (shown as the diagonal cylinder at the lower right) isoperating in air in a well-illuminated room at normal videodisplay brightness (∼100 cd/m2) (from ref 396).

1880 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 89: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

same route as discussed for TFTs on Si substrates.Transistor characteristics, such as those reported

for R-6T-based channels (see Figure 5-23b) are bro-ken into two operating regions: the linear regime atlow VDS, and the saturation regime at higher VDS.Highest transistor gain is achieved in the saturationregime. The drain current for an ideal transistor isthus313

Here, εi is the permittivity of the gate insulator, d isits thickness, VT is the threshold voltage, µeff, is the“effective” carrier mobility, and IDSS is the drainsaturation current. The voltage gain in a commonsource circuit configuration, where RL is the loadresistance at the TFT drain, is Av ≈ gmRL, where gmis the transconductance (i.e., gm ) [dIDS/dVGS]). Sincegm is largest in saturation, we obtain

Now the gain of the amplifier is also limited by theoutput conductance of the TFT, go, which is simplythe slope of the drain current characteristics insaturation, viz.: go ) [dIDS/dVDS]. The maximum

gain is therefore Av < gm/go, suggesting that practicaltransistors require a minimum output conductanceand a maximum transconductance. Furthermore, alow go assures that the channel leakage current inthe “off” state (IDO) will be small. Coupled with alarge mobility which ensures a large IDSS, the switch-ing ratio (IDSS/IDO) is large. Indeed, IDSS/IDO > 106 isnecessary for most high contrast AMLCD applica-tions. Finally, the frequency response of the transis-tor is determined by the cut-off frequency:

where CGS is the gate-source capacitance. From theforgoing discussion, it is apparent that the criticalperformance parameters of a TFT, i.e., its maximumcurrent, IDSS, transconductance, gm, and operationalbandwidth,∼fT, all depend on the figure of merit, µeff/L, which must be maximized, along with minimizinggo to ensure high-gain operation. In addition, it isimportant that VT be small (approaching 0 V) to allowfor low-voltage operation. Since L is a function ofdevice geometry, it is important to engineer struc-tures where L ≈ 2-5 µm can be routinely achieved.Both go and µeff are determined by thin-film proper-ties, with the optimal values achieved using highlyordered and high purity organic thin films such asthose grown by OMBD. Since µeff is primarily dueto interface properties between the channel layer andgate insulator,267 go can be reduced by maintaininga highly flat, but thin channel film in order todecrease current shunts through the film bulk. Filmflatness is routinely achieved by OMBD growth,which is probably essential for routine fabrication ofhigh performance TFTs competitive with existinghydrogenated amorphous Si (a-Si:H) transistors.As inferred from these considerations, the channel

mobility (µeff) is not generally the same as its bulk,or intrinsic, value (µo). Thus, µeff is determinedprimarily by defects at the interface, which maygenerate deep trapping levels. Alternatively, suchdefects can arise in the film itself due to boundariesbetween crystalline domains in the channel.267,268,402,403In any case, µeff, which is inferred from the draincurrent characteristics of the TFT (see eq 5.7), is often1-3 orders of magnitude less than µo. This is clearlyindicated from the data in Figure 5-24, where µeff is

Figure 5-23. (a) Schematic view of an organic thin filmtransistor fabricated on a Si substrate. W and L is thechannel width and length, respectively, Vg and Vd is thegate and drain voltages, and Id is the drain current. (b)TFT drain current characteristics. The TFT structureconsists of a R-6T organic channel layer deposited betweenAu source and drain electrodes on a glass substrate. Theinsulating gate layer is polymethacrylate with a capaci-tance, Ci ) 10 nF. Device dimensions are W ) 5 mm, L )50 µm (from ref 268).

IDS ) (W/Ld)εiµeff[(VGS - VT)2 - VDS

2/2]VDS < VGS - VT (5.7a)

IDSS ) (W/2Ld)εiµeff(VGS - VT)2 VDS > VGS - VT

(5.7b)

gm ) (W/Ld)εiµeff(VGS - VT) (5.8)

Figure 5-24. Mobility as a function of electric field (EDS)in the channel. Gate dimensions are L ) 25 µm (full dots),L ) 12 µm (open squares), L ) 4 µm (full triangles), and L) 1.5 µm (open triangles). The bars indicate the error inthe measurement of the mobility which have a slightdependence on VGS (from ref 403).

fT ) gm/2πCGS ) (µeff/2πL2)(VGS - VT) (5.9)

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1881

Page 90: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

plotted vs electric field for R-6T TFTs with differentchannel lengths. It is found that µeff < 0.03 cm2/(Vs) for the shortest gates (and hence those gates whosedimensions are comparable to the crystalline grainsize), a value considerably smaller than expected forbulk material (indicated by the dashed line). Forthese reasons, devices employing gate insulators havecharacteristics significantly inferior (i.e., lower gains,higher operating voltages, larger dark channel con-ductivities) to those based on thermally grown SiO2

on Si substrates.Attempts to increase µeff by doping, as has been

demonstrated for In and Sb incorporation into C60

channels,408 unfortunately also increases go to unac-ceptably high levels due to the resulting increase inchannel conductivity by nearly 6 orders of magnitude.The device whose characteristics are shown in

Figure 5-23b was fabricated on a poly(parabanic acid)resin substrate which allowed for flexing of thecompleted devices. The metal (Ag or Au) gateelectrode was sputter deposited onto the substrate,followed by the casting of a polymeric gate insulator(cyanoethylpullulan). Vacuum deposition of 500 Åof R-6T formed the channel, followed by evaporationof Au source and drain electrodes. The completeddevice had L ) 50 µm, W ) 2.5 mm. Deviceperformance characteristics include gm ) 3 × 10-7

S, µeff ) 0.46 cm2/(V s), VTH ) -1.3 V, go ) 10-8 S atVDS ) -20 V and VGS ) -8 V, and fT ) 30 kHz. Thehighest µeff reported269,404,409 is from 0.7 to 1.5 cm2/(Vs) obtained for vacuum-deposited pentacene films onthermally grown SiO2 using the TFT configurationshown in Figure 5-23a. Similar structures have alsobeen demonstrated267 with IDSS/IDO > 106. Thesevalues are similar to those obtained313 with a-Si:Hused in AMLCDs. While a weak increase in mobilityof many organic films has been observed with in-creasing electric field61,267,381,410 (and therefore VGS),there is little prospect for exceeding µeff ≈ 1-5 cm2/(V s) at room temperature using even perfectlyordered planar stacking molecules since the µo isultimately limited to these values by thermallyinduced molecular librations. For pentacene TFTs,µeff ≈ µo.Given µeff ) 1 cm2/(V s), L ) 5 µm, and VGS - VT )

10 V, we obtain a maximum bandwidth of fT ) 8 MHzwhich is suitable for many display and low-costelectronic circuit applications.The role of mesoscopic organization has been

investigated for 6T-based TFTs, where the substratetemperature was varied during vacuum deposition.411For this molecular structure, charge propagates alongthe stacking axis which therefore must be orientedparallel to the substrate plane to ensure high-channelmobility using the geometry in Figure 5-23a. Devicesfabricated by deposition at room temperature orbelow tend to nucleate with their stacking axisperpendicular to the plane,412 thus leading to a lowmobility (2 × 10-3 cm2/(V s)). The mobility is furtherreduced for growth at the lowest temperatures (77K) due to impurity adsorption onto the cold sub-strates. As noted in section 2, deposition onto cooledsubstrates necessitates growth under UHV condi-tions if impurity incorporation into the film is to be

minimized. The room temperature deposition re-sulted in polycrystallites with ∼50 nm diameters.However, at higher temperatures (g190 °C) thepreferred molecular orientation with stacking paral-lel to the substrate plane is achieved. Large poly-crystallites (30 × 200 nm2) with this orientation areobserved, leading to mobilities as high as 2.5 × 10-2

cm2/(V s) along with a low film dark conductivity. Atthe highest temperatures of growth (260 °C), roughsurface morphology develops, typical of most of thematerials discussed in section 3. While these filmsdo not have a mobility as high as pentacene, theynevertheless have a very low dark conductance(∼10-7 S cm-1) required for high transistor gain andswitching ratio. Attempts to anneal the as-grownlayers of lutetium and thulium bisphthalocyanines(Pc2Lu and Pc2Tm, respectively), as well as for R-6T,at 150 °C has also led to an increased domain size,and hence an increased mobility.267,407 As discussedin section 3, this high-temperature process can resultin some film rearrangement which may increase thepolycrystallite diameter in the channel. Systematicstudies of rapid thermal annealing of R-6T TFTs haveshown413 that heating the film near to its meltingpoint significantly increases grain size, which leadsto a concomitant increase in the device on/off (IDSS/IDO) ratio and a decrease in VT to near 0 V. Indeed,heating above the film melting point appears tocompletely eliminate deep grain boundaries betweenadjacent R-6T islands.Since most TFTs are generally operated in the

accumulation mode, this places constraints on thetypes of circuits that can be realized. The mostubiquitously employed integrated circuit technologyused today is based on Si complementary metal-oxide semiconductor (CMOS) field effect transistor(FET) electronics. The ability to fabricate both p- andn-channel FETs in a single integrated circuit plat-form allows for the realization of very low powerdissipation circuits which only draw current when atransistor is switched. However, in circuits whereonly a single layer organic film is used for the channelof all transistors, complementary logic is not possible.Dodabalapur et al.266,414 addressed this issue bydemonstrating a novel device in which a 100-200 Åthick p-channel layer (R-6T) and a 200-400 Å thickn-channel layer (C60) were sequentially deposited onan SiO2 gate (L ) 4 µm) insulator. By using eitherpositive or negative VGS, the transistor was able toconduct either electrons or holes in the accumulationmode. The operation of this device can be understoodin terms of the band diagram in Figure 5-25a, whichshows the results of ionization potential measure-ments of the relative positions of the HOMO andLUMO levels of R-6T and C60 obtained in vacuo fromphotoelectron spectroscopy data.415,416 When VGS >0, the HOMO of the R-6T rises above the Fermienergy (left diagram, Figure 5-25a), opening conduc-tivity in the p channel which is in the accumulationmode. However, when VGS < 0, the energy “well”formed at the R-6T/C60 heterointerface (right dia-gram, Figure 5-24a) provides an n-type accumulationchannel. The resulting “bipolar” IDS vs VDS charac-teristics in both n and p accumulation modes are

1882 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 91: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

shown in Figure 5-25b, with a clear shift in polarityof IDS and VDS for these two cases. Analysis of thesecharacteristics in both modes yields an effectivechannel mobility of µeff )4 × 10-3 cm2/(V s) (with VT≈ 0 V) for p-channel operation, and µeff ) 5 × 10-3

cm2/(V s) (with VT ≈ 40 V) for n-channel operation.While both the mobility and threshold voltages of thisTFT were not as good as discrete p- and n-channelTFTs, it nevertheless is an important demonstrationof the potential for using organics for complementarylogic circuit applications. Indeed, these device char-acteristics critically depend on the film thicknessesused, as well as on their crystalline quality,266 sug-gesting that there remains considerable room forimprovement as these parameters are brought undercontrol. A further outcome of this work is that itdemonstrates that semiconductor-like band picturescan be used to understand some of the more complexbehavior of narrow bandwidth molecular organiccompounds employed in electronic device structures.

6.0. The Future of Ordered Organic Thin FilmsGrown in the Vapor Phase

6.1. Limitations of OMBDIn this review, we have shown that vacuum growth

of small molecular weight organic molecular crystalscan lead to unprecedented control over the structureand purity of the resulting thin films. This isparticularly true when using the ultrahigh-vacuumprocess of organic molecular beam deposition em-ploying highly purified organic source materials.OMBD has been demonstrated to result in long rangecrystalline order of the as-grown films. If the ap-propriate thermodynamic conditions are chosen dur-ing growth, unique crystalline thin-film structurescan be realized whose characteristics depend on thedetailed balance between growth kinetics, substrate-film interactions, and the degrees of freedom avail-able to the organic crystal itself. Long-range orderand strain incorporated during growth do not alwaysresult in defects. On the contrary, “templating” bythe substrate has led to the growth of crystallineisomorphs which differ significantly in both structureand properties from the bulk. Overall, lattice match-ing of organic films to the substrate structure doesnot appear to be influential in generating low defectdensity films, as is the case in inorganic semiconduc-tor epitaxial film growth.By manipulating the film structure and achieving

ultrathin, ordered OMCs, OMBD has made thegrowth of multilayer structures such as organicmultiple quantum wells a reality. These organicMQWs have been instrumental in revealing impor-tant physical properties of OMCs which were, priorto this point, inaccessible. For example, organicMQW structures have been grown which clearlydemonstrate excitonic quantum size effects, provid-ing, for the first time, direct measurements of excitonradius of extended charge-transfer states which havelong been viewed as an essential element in organicsemiconductor photoconductivity and other optoelec-tronic phenomena. The film perfection achieved inOMBD growth has led to increased carrier diffusionlengths which has already been used to increaseorganic solar cell efficiency, as well as to increase theeffective channel mobility in organic thin-film tran-sistors. In particular, initial results obtained forMQWs consisting of organic heterojunctions as wellas hybrid organic/inorganic multilayers have sug-gested that these structures may provide significantperformance improvements in photovoltaic and pho-toconductive devices. Perhaps the most complexdevice demonstrated to date employing OMBD-grownfilms, and one that most clearly demonstrates theirpromise, is the three-color, stacked light emittingdiode discussed in section 5.3 (see Figure 5-21). Thatdevice, consisting of 12 different layers composed ofamorphous and crystalline organic thin films, metalsand conducting oxides, to our knowledge can only beachieved using vacuum-deposited organic semicon-ductor materials. The successful demonstration ofsuch a complex structure was made possible bysteady scientific and engineering progress made inunderstanding organic molecular compounds by nu-merous laboratories worldwide over the previous 10-15 years.

Figure 5-25. (a) Proposed energy band diagrams of theheterojunction transistor in the p-channel and n-channelmodes of operation. In the p-channel mode, an accumula-tion layer of holes is formed at the R-6T/SiO2 interface,while in the n-channel mode, the accumulation layer ofelectrons is formed in the C60 at the C60/R-6T interface. (b)Drain current characteristics for the heterostructure TFTin p-channel mode (A) and n-channel mode (B) of operation.The device gate length is L ) 4 µm (from ref 266).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1883

Page 92: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

In spite of this remarkable progress, vacuumgrowth of highly ordered organic films is only justnow being exploited to the extent that practicalapplications may be achieved for active optoelectronicorganic thin-film devices such as those discussed insection 5. There still remain both significant op-portunities and problems in OMBD growth of organicthin films and the devices which employ thesematerials. For example, many of the same capabili-ties provided by MBE of inorganic compounds mayalso apply to the growth of organic thin films. Weshowed in section 3.4 that molecular layer deposition(which is analogous to atomic layer epitaxy) can beused to obtain saturated monolayer growth controlover organic superlattices.50 This was implementedby precise substrate temperature control which al-lowed for selective adsorption of alternating mono-layers of functionalized organic molecules. Extend-ing this control to two dimensions via such techniquesas laser-assisted growth417 may eventually allow forin situ patterning of films across a substrate surface.Electric fields applied either parallel or perpen-

dicular to the substrate surface during growth inUHV has also been demonstrated to be useful inmanipulation of structure. These techniques havebeen applied to both small molecular weight polar418and nonpolar419 OMCs, and polymers420-424 as ameans to orient dipolar molecules with large nonlin-ear optical coefficients. The applied fields are ameans for in situ molecular poling, similar to meth-ods used to align electrooptic chromophores in poly-mers heated above their glass transition tempera-ture.14 For this purpose, orienting the field in-planeusing closely spaced electrodes pre-patterned on thesubstrate prior to growth419-421,423,424 can be moreeffective in generating high fields which are suf-ficient(>105 V/cm) to induce molecular ordering, thanusing fields oriented perpendicular to the substrateusing external field apparatus such as grids,418,422where comparatively small fields (∼104 V/cm) arepossible.The use of electric fields in orienting molecules

during growth was clearly demonstrated in twoexperiments performed on strongly polar,425 vi-nylidene flouride (VDF)-based thin films. In the firstseries of investigations,422 3000 Å thick VDF/trifluo-rethylene (TrFE) copolymer films with 73% VDFwere deposited in vacuum at a pressure of 3 × 10-6

Torr at a rate of 2.5 Å/s and with Tsub ) 25 °C. Inthis case, electric fields as high as 62 kV/cm wereoriented normal to the substrate which consisted ofa metal (Cu or Al) contact predeposited onto anoptically flat SiO2 slide. The orientational depen-dence of the resulting films on electric field are shownby the energy dispersive X-ray diffraction patternshown in Figure 6-1a. Here, the peak identified asthe (110) or (200) Bragg reflection is seen to increasein strength as the field is increased. This feature isdue to the molecular axes of form I VDF/TrFEcrystals oriented parallel to the substrate as field isincreased. This is equivalent to the molecular dipolemoment, shown schematically in Figure 6-1b alsolying parallel to the electric field direction. Asanticipated, the degree of alignment was reduced as

the substrate was heated, resulting in nearly com-plete disorder as the temperature was increasedabove 100 °C.As a final test of the usefulness of the application

of the electric field during the growth process, thestructure of films grown under high electric field werecompared with those grown in the absence of suchfields. It was found that only a very small (110),(200)peak intensity was observed for the latter films ascompared with that observed for films depositedunder high electric fields (42 kV/cm). In addition, itwas not possible to recover the peak intensity by post-poling the precrystallized films; that is even after anelectric field was applied in vacuuo at Tsub ) 140 °Cfor 1 h immediately following deposition, no ad-ditional film alignment was induced.In a second series of experiments on a similar

compound, the effectiveness of in-plane fields wasexplored as a means for molecular “poling” duringgrowth.423 In this case, deposition of poly-VDF(PVDF) was done on glass substrates in the 50 µmgap between Al electrodes across which fields of upto 160 kV/cm were applied. Conditions for theseexperiments included deposition at 7.5 × 10-6 Torrat a rate from 0.18 to 0.31 Å/s for substrate temper-atures maintained between -130 and 100 °C. As hasbeen widely observed for thin film growth by OMBD(see section 3), rough surfaces were observed at thehighest (∼100 °C) temperatures, and thus onlyresults for Tsub < 80 °C were reported. Second-harmonic intensity (SHG) of the as-deposited filmswas used as the means to investigate the order of

Figure 6-1. (a) Total reflection dispersive X-ray diffractionpatterns of evaporated films of VDF/TrFE showing thedependence of the (110) and (200) reflections for variousapplied electric fields. (b) Schematic representation of theproposed molecular orientation to the substrate when thesample is evaporated with (left) and without (right) theapplication of an electric field (from ref 422).

1884 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 93: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the product. In this case, a decreasing ratio of thenonlinear susceptibility tensor elements, R ) øxyy/øxxxis interpreted as an increase of dipole orientation inthe direction of the applied electric field (i.e., in thesubstrate plane defined as the x,y plane in this work).Two observations were made, as is apparent inFigure 6-2: (i) Dipole alignment increases as sub-strate temperature decreases, as is apparent from adecrease in R by a factor of two as Tsub is loweredfrom 100 to -130 °C. (ii) The SHG efficiency of filmsdeposited in the absence of an electric field, and thenpost-poled in situ at Tsub ) 25 °C in an electric fieldof 160 kV/cm was negligibly small compared to filmsdeposited in the presence of such a field (Figure 6-2b).This latter result is similar to that found by Yoshidaand co-workers422 in applying fields normal to thesubstrate. This suggests that the dipoles in post-poled films could not be efficiently reordered sincethe energy stored in the large crystalline domainswhich are produced during growth (particularly whenTsub is small) is very large. In contrast, duringgrowth, the molecules themselves along with smallclusters of molecules nucleating on the surface, areeasily reoriented by the application of the modestelectric fields employed in these experiments.These experiments are very promising in that they

open new avenues for manipulating the structure ofpolar molecular crystals, thereby enhancing a desiredproperty of the materials. Many other methods, suchas application of magnetic fields, local substratetemperature control, or irradiation with strong opti-cal fields may also prove useful for in situ tailoringthe structure of OMCs deposited by OMBD in thenear future.One significant shortcoming of OMBD is that it is

inherently limited to growth of compounds which areUHV compatible. While this property is essential ifenvironmentally stable, single-component systemsare desirable, which is overwhelmingly the case foruse in any practical application, there are numerouscompounds consisting of two or more organic molec-ular species whose structures might individually bevolatile (and hence incompatible with UHV deposi-tion processes), although the compound itself might

be highly stable. An example compound with theseproperties is the organic salt, 4′-(dimethylamino)-N-methyl-4-stilbazolium tosylate (DAST) shown in Fig-ure 6-3a, which has been shown to have a second-order susceptibility value of ø(2) ∼ 103 times greaterthan that of urea due to dipole alignment of its cationand anion constituents.426 Although in situ reactionsof multicomponent organic molecules to synthesizepolymer films previously have been demonstratedusing such vacuum techniques as physical vapordeposition or vapor deposition polymerization,427 at-tempts in our laboratory at double-source coevapo-ration of the DAST neutral precursors 4′-(dimethy-lamino)-4-stilbazole (DAS) and methyl p-toluene-sulfonate (methyl tosylate, MT) to form DAST havebeen unsuccessful due to the radically different vaporpressures of DAS and MT, which leads to highlynonstoichiometric growth. To solve problems ofincongruent growth employing high-vapor pressuremolecular precursors incompatible with the UHVenvironment, vapor-phase epitaxy has been devel-oped to grow epitaxial thin films of many III-Vcompound semiconductors, such as InP and GaAs.428Here, a high vapor pressure compound (typically ametal halide or a metallorganic) of each respectivemetal is carried independently, via a carrier gas, intoa high-temperature reaction zone. In that zone, thecompounds are deposited onto a heated substratewhere they thermally decompose and react to yield

Figure 6-2. (a) Ratio of second-order nonlinear suscepti-bility tensor elements, R ) øxyy/øxxx for PVDF films as afunction of substrate temperature during growth. As Rdecreases, the dipole orientation in the direction of theapplied electric field (i.e., in the substrate plane) increases.(b) Relative second-harmonic generation intensity as afunction of annealing temperature for films deposited inthe presence of an electric field (open circles), and for filmsdeposited in the absence of field and poled after deposition.The SH intensities are normalized to their room temper-ature values (from ref 423).

Figure 6-3. (a) Perspective of a unit cell of DAST and (b)proposed scheme for the synthesis of DAST molecules byorganic vapor phase deposition (from ref 430).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1885

Page 94: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

the desired III-V compound. The excess reactantsand reaction products are then exhausted from thesystem via a scrubber.An analogous technique was recently developed to

grow DAST by the reaction of two volatile organicmaterials in a hot-wall, atmospheric pressure reac-tor.429,430 Nuclear magnetic resonance (NMR) analy-sis was used to determine that the stoichiometry ofpolycrystalline DAST films grown by this noveltechnique was >95% pure (limited by instrumentalsensitivity). By using X-ray and e-beam diffractionand other analytical methods, a significant depen-dence of film quality, such as ordering and crystallitesize, on the substrate composition and other deposi-tion conditions used for growth was found. Thissuggests that it may be possible to generate opticalquality thin films of DAST and similar organic saltsand compounds by OVPD using suitable substrates.The reaction scheme for the preparation of DAST

by OVPD is shown in Figure 6-3b. In step 1, thermaldemethylation of 4′-(dimethylamino)-N-methyl-4-stil-bazolium iodide (DASI) gives neutral DAS with theelimination of volatile CH3I. In step 2, the DASreacts with MT, resulting in methylation of thepyridine ring at the nitrogen atom. The resultingstilbazolium and tosylate ions form DAST crystals.The OVPD reactor designed to grow DAST by the

proposed scheme is shown schematically in Figure6-4. The reaction tube is a hot-walled glass cylinderin which an alumina crucible filled with a smallamount of DASI is placed in a hot zone of the furnacewhile a stream of pure N2 gas provides a slightpressure gradient, causing the DAS vapor to flowalong the reactor tube. The very high vapor pressureMT is loaded into a glass bubbler whose temperature

is maintained by a silicone oil bath. Again, N2 gaspasses through a flow meter and bubbles through theMT, thereby carrying its vapor through a heatedtransfer tube (to prevent recondensation prior tointroduction into the hot zone) into the furnace tube.The DAS vapor mixes with the MT vapor in thereaction zone, along which a temperature gradientis maintained. There it reacts to form DAST onsubstrates supported along the length of the tube.Excess, unreacted MT vapor, and any volatile side-reaction products, are exhausted from the cold endof the reactor. To avoid diffusive flow conditionswhich tend to result in highly nonuniform filmmorphology, growth is carried out at reduced pres-sure (typically 5 Torr).Optimal growth of the reddish-brown DAST films

results, thus far, in somewhat textured film surfacesas shown in the micrograph in Figure 6-5, where asurface roughness of ∼200 Å is observed for the 2µm thick film shown. Such films are obtained atsubstrate temperatures between 140 and 150 °C inthe ∼200 °C reaction zone. The DASI crucibletemperature is 253 ( 2 °C, and the MT bubbler is 55( 2 °C,430 resulting in growth rates of ∼1-10 Å/s,depending on the gas flow rates and the diameter ofthe reactor tube. The films, grown on amorphous Ausubstrates, are polycrystalline,429 suggesting thatsubstrates with appropriate structure and symmetryare needed to initiate more ordered growth. Never-theless, the structural and chemical purity of thefilms give rise to intense second harmonic generationat λ ) 0.95 µm, approaching that of pure, orientedDAST crystals.429

While the OVPD method has shown promise ofgrowing multicomponent films consisting of organic

Figure 6-4. Schematic layout of a low-pressure, organic vapor-phase deposition system currently used in the author’slaboratory.

1886 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 95: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

molecular constituents with widely different vaporpressures, as in the case of most such techniques, itis difficult to maintain precise control over bothsubstrate and reactor temperature. This suggeststhat adapting the method to growth in UHV condi-tions has the potential of yielding very ordered films(similar to those discussed in section 3) while at thesame time assuring the purity of the product. Sucha technique would be similar to “chemical beamepitaxy”, or CBE, employed in the growth of III-Vand II-VI compound semiconductor materials.158,431,432By analogy to CBE, the high volatility molecularconstituents can be introduced into a UHV, organicchemical beam deposition (or OCBD) chamber similarto that shown in Figure 2-1 via a bubbler andappropriate gas injector. The low vapor pressureadducts can be evaporated from standard effusioncells, similar to those used in conventional OMBD.By heating the precursor molecules to sufficientlyhigh temperatures, they can be reduced to their ionicstates (i.e., “cracked”), thereby allowing for thegrowth of the desired, two-component compound.This straightforward hybrid combination of OMBDand OVPD may eventually allow for the growth of aconsiderably expanded range of organic materials,including organic salts (such as nonlinear opticalcompounds exemplified by DAST, superconductingmaterials such as the Bechgaard salts, or charge-transfer complexes), and doped compounds such asthose required for high luminance organic lightemitting devices (e.g., TPP in Alq3).Note that in the example of guest-host systems,

controlling the doping of one molecular species at lowconcentrations (∼1%) in a high concentration organicmatrix, is extremely difficult using conventionalvacuum evaporation techniques. The reason for thisdifficulty is easily understood when we considerthat the evaporation rate follows an exponentialdependence: R ≈ exp(-Ls-g/kT), where Ls-g is thelatent heat of sublimation of a material. To introducea small concentration of a guest chromophore into ahost matrix, the temperature of the Knudsen cell

must therefore be very carefully adjusted to avoidrapid, uncontrolled evaporation. However, using theprocess of OVPD (or OCBD), the dilution of the guestis primarily determined by the carrier gas mass flowrate rather than the bubbler or boat temperature,which is far easier to control to the required accuracy.Hence, OVPD, or its high-vacuum analog, OCBD,potentially offers many advantages in the depositionof organic thin-film materials consisting of more thana single molecular organic component.Finally, we note that the ability to initiate ordered

growth on all substrates requires the existence of asubstrate potential which is spatially varying witha symmetry similar to that of the adsorbed film.While we have seen that such substrate “templating”may exist on glass substrates,65 polymers,102,433,434step edges,115 rubbed substrates,179 and even crys-tals,70,73,98,111 the requirements which must be satis-fied to provide a substrate leading to a high degreeof order in the resulting film are not straightforwardto achieve. One very promising approach which hasonly just begun to be followed is that of substrate“templating”, or derivatization, using self-assembledmonolayers of functional organic molecules prede-posited (generally from solution) onto the inorganicsubstrate surface.4,435,436 A surface molecule whichchemically reacts with the vacuum deposited speciescan anchor the deposited molecule in an orderedmonolayer as it arrives at the substrate. To besuccessful, the functionalized molecules must bedeposited (perhaps by Langmuir-Blodgett or otherwet chemical process) in such a way that a saturatedmonolayer completely fills the substrate surface areaprior to the vacuum deposition of the desired thinfilm molecular species. These self-assembly tech-niques provide a unique ability to achieve orderingof the resulting layers. Properties that this templat-ing layer must possess are compatibility with thesubstrate temperatures, ultrahigh-vacuum environ-ment, and chemistry of the material to be depositedin the OMBD chamber.In addition to our need to extend the capabilities

of the growth technology as discussed above, thereremain numerous open questions regarding thephysics and chemistry of OMCs. Given the struc-tural and material purity afforded by this precisiongrowth technology, we now have, for the first time,a clear opportunity to construct structures whichprobe the fundamental physics of this fascinatingclass of materials. A primary goal of past work onUHV growth of organic thin films was to understand,in detail, the conditions which lead to organizedmesoscopic and macroscopic molecular order. Whileinitial work was focused on understanding equilib-rium structure, almost nothing has been done toexplore the thermodynamic conditions leading tonear-perfect thin-film growth. Acquiring such anunderstanding should have the immediate conse-quence of providing insight into the class of materialsand the growth conditions needed to achieve long-range crystalline order. The nature of extendedexciton states appears to be accessible using OMBD-grown organic nanostructures. Correlation effects,the dependence on orbital overlap both in-plane andnormal to the stacking axis, charge transport acrossorganic and organic/inorganic interfaces, and many

Figure 6-5. Optical micrograph of a 2 µm thick film ofDAST grown by organic vapor phase deposition. A slightsurface texture is apparent from this micrograph. Thesecond-harmonic generation intensity of such films isapproximately equal to that obtained for bulk DAST grownfrom solution (ref 429).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1887

Page 96: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

other fundamental phenomena still require consider-able study using such structures before this class ofmaterials can be fully understood and exploited fora particular device application. We note that manyorganic systems such as planar stacking molecules,typified by PTCDA, support one-dimensional trans-port. By growing ultrathin films of such materials,a 0D environment is created. Hence, OMBD ofnearly perfect nanostructures appears to be an excel-lent means to probe transitions from 3D to 2D to 1Dto 0D phenomena.Finally, fundamental questions persist on the role

that defects play in determining the optoelectronicproperties of OMCs. We still need to explore howdislocations, molecular fractions, and other impuri-ties affect carrier transport and the optical propertiesof this important and enormous class of materials.Once this is fully understood from both a theoreticalas well as empirical viewpoint, we will be able toappreciate the importance of UHV growth techniquessuch as OMBD in achieving and tailoring the materi-als properties desired.

6.2. Applications on the HorizonIt is becoming increasingly apparent that the

control over thin-film structure (with the ability tomonitor the evolution of structure in real time), andultrahigh materials purity uniquely afforded bygrowth in UHV are important for producing highperformance devices with manufacturing yields andlong operational lifetimes sufficient to generate acost-effective device technology. The properties ofalmost all optical and electronic devices are sensitiveto control of vertical device dimensions (i.e., layerthicknesses) and material doping. For example, apractical display technology must be based on pixelswith highly uniform operating characteristics (e.g.,operating voltage and current), and extremely highdevice yields across the full display area which mayconsist of >106 individual light-emitting elementsextending over >103 cm2. As we have shown insection 5.3, these properties are strongly determinedby the ability to deposit very thin films (∼200-500Å thick) with extremely high thickness uniformityacross these large areas. To our knowledge, onlyvacuum deposition currently provides this level ofcontrol, and only UHV processes such as OMBD canensure the requisite control over material purity.A key performance factor of any practical technol-

ogy is cost. Indeed, the potential for producing anentirely new class of extremely low cost devices isone of the principal reasons that organic thin filmsare of interest. Hence, it is reasonable to questionwhether UHV techniques can lead to low productioncosts. Experience in III-V-based devices has cer-tainly indicated that MBE growth can produce largevolumes of ultralow-cost components, the most no-table example being the fact that in 1995, ∼50% ofthe worldwide production of compact disc lasers437(with a wholesale price of ∼$1) was accomplished viaMBE. Hence, while the infrastructural costs of UHVgrowth are high, this does not necessarily present abarrier to the low-cost manufacture of optoelectroniccomponents. The purity and uniformity afforded byOMBD appear to be essential for the understanding

of organic materials, and in particular the growth ofsolar cells, OLEDs, TFTs, and numerous other de-vices. Whether or not these conditions can be relaxedin an effort to further reduce cost will largely dependon the requirements of the specific application.Numerous device opportunities which can be

addressed using vacuum deposited organic thinfilms remain to exploited. One area of particularinterest is in electrically pumped, organic thin-filmlasers.390,438-443 Such devices have potential applica-tion for very low cost optical memory reading andrecording, printing, laser scanners, etc. Some uniquefeatures of electrically pumped organic lasers wouldinclude ease of integration with electronic intercon-nection ribbons and other such “platforms”, acces-sibility to a wide range of emission wavelengths viamolecular compositional tuning, and insensitivity ofthe output power and wavelength to wide variationsin operating temperature.Stimulated emission by optical pumping of solid-

state organic materials is well known, with the firstdemonstration of lasing in dye-doped gels and mo-lecular crystals in the late 1960s.444-447 Demonstra-tion of similar effects in conducting organic thin filmscan therefore lead to an entirely new class of electri-cally pumped laser diodes based on organic semicon-ductor thin films, provided that sufficiently lowoptical pump threshold energies can be obtained.Lasing (in contrast to amplified spontaneous emis-

sion, superluminescence or related phenomena) canbe unambiguously identified from five phenomena:(i) a clear indication of a threshold in output energyas a function of input (or pump) energy, with a highlasing efficiency above threshold, (ii) strong outputbeam polarization, (iii) spatial coherence (as indicatedby a diffraction-limited output beam or speckle), (iv)significant spectral line narrowing, and (v) the exist-ence of laser cavity resonances, or modes. Whileseveral recent reports on polymer-based opticallypumped thin films have discussed structures exhibit-ing one or two of the above phenomena,439-441,448 toour knowledge, the only report to date where lasingis clearly identified in conducting organic films is inAlq3 doped with DCM443 (see Figure 6-4). In thatwork, a very low lasing threshold at a pump energydensity of 1 µJ/cm2 with a 500 ps excitation pulse,was reported.Alq3 doped with DCM provides an excellent laser

material443 since the red stimulated emission of DCMat a wavelength of λ ) 645 nm is far from theultraviolet absorption edge of the Alq3 host (at λ )450 nm),347 while the absorbance of the DCM iscentered near the emission maximum of Alq3 (at λ )530 nm), thereby providing for efficient Forsterenergy transfer196 from UV-excited Alq3. Doping alsoallows for reduction of the density of the opticallyactive DCMmolecules (thereby reducing the effectivedensity of states), which lowers the threshold andincreases the efficiency of a laser.449 In fact, efficientorange electroluminescence for DCM doped Alq3 filmshas been demonstrated.264

Lasers shown in Figure 6-6 were grown on InPsubstrates precoated with a 2 µm thick layer of SiO2,deposited by plasma-enhanced chemical vapor depo-sition. The organic films were deposited by high-

1888 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 97: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

vacuum coevaporation of 40:1 (mass ratio) Alq3 andDCM. A 3000 Å thick film of Alq3/DCM (with opticalindex of refraction n ) 1.7) forms a slab opticalwaveguide, with SiO2 (n ) 1.4) as a cladding layeron one side, and air (n ) 1) on the other. Prior toorganic film deposition, the (100) InP substrate wascleaved, ultimately defining the laser length whileforming parallel edges. The vacuum-deposited or-ganic film conforms to the shape of the underlyingsubstrate, resulting in optically smooth, parallelfacets with reflectivities ∼7%.The pump provided lateral confinement using the

focused beam from a pulsed nitrogen laser (λ ) 337nm, pulse width ) 500 ps at a 50 Hz repetition rate),as shown in Figure 6-6. A photograph of an operat-ing slab laser device also shown in Figure 6-6indicates a well-defined beam of bright red laseremission in a direction orthogonal to the facets.Figure 6-7a shows a high-resolution spectrum for ashort cavity, 500 µm long slab laser, revealingnumerous peaks (inset) evenly spaced at 2 Å and eachwith a <1 Å FWHM, corresponding to the freespectral range of the cavity, clearly indicating reso-nant cavity modes. The laser output power washighly stablesi.e., no significant degradation in thelaser performance was observed after operating ∼24h at a very high peak power of ∼50 W in a nitrogenatmosphere (∼4 × 106 pulses). Such stability isessential for the eventual practical application ofelectrical injection lasers based on these compounds.Finally, the dependence of output energy on thepump energy for the devices, shown in Figure 6-7b,clearly indicates a threshold. To reduce mirrorlosses, long (∼1 mm) laser cavities were used toobtain the differential quantum efficiency (η) abovethreshold. It was found that η ) 30% for the slabdevice.Several advantages to using vacuum deposited

organic films are readily apparent from these earlyexperiments. For example, facet formation is easily

accomplished simply by depositing the films on asubstrate with sharp, parallel edges. Alternatively,the facets can be formed by cleaving of the substrateafter film deposition. In either case, no externalfeedback, or etched facets which tend to reduceperformance are required, as may be the case inpolymer structures. Furthermore, the films are

Figure 6-6. (a) Schematic of the laser structure and experimental setup, (b) photograph of the Alq3/DCM slab waveguidelaser showing the red output beam diverging with distance from the laser facets indicated by the intense radiation of theblack InP substrate (photo courtesy of A. Forrest using the Princeton shield for contrast), and (c) chemical structuralformulae of DCM and Alq3 (from ref 443).

Figure 6-7. (a) High-resolution edge emission spectrumof a 500 µm long slab laser at a pump energy density of1.7 µJ/cm2. Inset: A portion of the same spectrum underhigher resolution, showing evenly spaced resonant cavitymodes and indicating a modal FWHM < 1 Å. (b) Depen-dence of output energy near threshold for a 1 mm long slabdevice. Solid lines are fits to the data yield an externalefficiency above theshold of ∼30% (from ref 443).

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1889

Page 98: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

extremely flat and smooth, resulting in very lowscattering losses along the cavity, along with theability to form very abrupt interfaces between ul-trathin organic layers which may be necessary in thefabrication of double heterostructure injection lasers.Compatibility between two different vacuum-depos-ited layers is also rarely an issue in choosing materi-als and structures best suited for a particular deviceapplication. Given that a laser typically consists ofseveral layers with different optical and electronicfunctions, this materials compatibility is an ex-tremely advantageous feature as compared to thesituation in polymers and inorganic semiconductors.Many issues must still be addressed before a

practical, electrically pumped device can be realized.Contact degradation presents the largest limitationto OLED lifetime. This problem is considerably moredifficult to resolve in lasers where the current densi-ties are increased by a factor of 105 or more over thatused in OLEDs. Indeed, if we assume that theintrinsic lifetime of the organic materials is ∼105 hat an intensity of 100 cd/m2 (which corresponds to adrive current of ∼1 mA/cm2 for most OLEDs ofmoderate efficiency), then the intrinsic CW solid stateorganic laser lifetime will only be in the range of1-10 h. This is clearly inadequate for most practicalapplications and must be increased by increasing theluminance efficiency (or optical gain) of the organicmaterials, by employing pulsed operation (which canincrease the lifetime by at least the inverse of thelaser duty cycle), and by choosing organic moleculeswhich are extremely robust in the presence of thelarge current densities and optical field intensitiesfound in laser cavities.Other devices which have the potential for trans-

forming various device industries include very highefficiency organic photovoltaics as discussed in sec-tion 5.2. Use of multilayer stacks to improve ef-ficiency has shown early promise, although to datethe best power efficiencies lie between 1% and 2%.The need for structural order and precision affordedby the OMBD process will be required to makefurther progress in this important application area.Whether or not success will ultimately be achievedwill depend critically on increasing the currentresearch effort concentrated in this field.There appear to be numerous opportunities for

organic thin-film devices in nonlinear optic applica-tions as well. However, most OMCs deposited invacuum possess a center of symmetry, eliminatingtheir usefulness in ø(2) applications such as electroop-tic modulators, etc. Once again, OMBD-grownMQWstructures may have some limited application here.By the appropriate choice of molecular species char-acterized by a very high degree of orbital overlapbetween adjacent molecules, such processes as thequantum confined Stark effect may be employed toincrease the second-order nonlinear effects in OMCs.A more promising approach, however, is to usecrystals with large built-in dipole moments such asthe organic salts. As discussed in section 6.1, suc-cessful growth of these multicomponent materialswill require OVPD or modified OMBD such asorganic chemical beam deposition. Furthermore, themodest vacuum combined with the horizontal glass

reactor geometry characteristic of OVPD and associ-ated techniques can be adapted to bulk materialsgrowth using roll-to-roll deposition processes. Thisapproach has considerable potential for drasticallyreducing the cost of growth of molecular organic filmsdeposited on polymeric or other lightweight, flexiblesubstrates.Finally, we note that significant challenges remain

in developing microfabrication processes for OMC-based devices. That is, a major impediment to thepractical application of these materials is their fragil-ity when exposed to water, solvents, elevated tem-peratures, and other conditions commonly encoun-tered in standard semiconductor fabrication.294 Hence,to fabricate high-resolution displays, TFTs, or opticalwaveguide devices requires redesign of many of theseprocessing methods. One approach is to employ alldry patterning using reactive ion etching and mul-tiple level photoresist and/or insulator (or metal)layers to prevent exposure to the wet chemistry usedin photolithographic processes.296 While early dem-onstrations of such modified fabrication methodshave achieved modest success, it remains to bedetermined whether or not employment of theseprocesses on a large scale will result in the highdevice yields and low costs required for practicalimplementation of small molecular weight organicthin-film devices.In conclusion, whether or not a new generation of

optoelectronic devices based on vacuum-depositedOMCs will gain a foothold in the spectrum of ap-plications currently being fulfilled by more conven-tional semiconductor materials depends on threefactors: (i) The ability of OMCs to provide a low-costsolution with higher performance than can otherwisebe achieved, (ii) the successful development of highyield, simple to implement, very low cost devicefabrication technologies, and (iii) the long-term op-erational and environmental stability of the materialsused in the devices. Enormous strides have beenmade in resolving all of these issues, and OMBD oforganic molecular thin films has played a major rolein improving the reproducibility, stability, perfor-mance and uniformity of organic nanostructures tolevels which only a few years were not thought to beachievable. Hence, it is very likely that this technol-ogy will soon lead to the realization of a large classof new optoelectronic systems based primarily onOMC thin films grown by high-vacuum deposition.Most probably, the first such demonstrations will beorganic electroluminescent displays which may be-come commercially available in the very near future.

7.0. ConclusionsIn this review, we have described in detail the

progress and current state-of-the-art in the high-vacuum growth technology of molecular organic thinfilms. We have shown that OMC thin films can begrown into highly ordered structures independent oflattice matching with the underlying substrate,provided that the substrate bond energy is suf-ficiently small. The resulting film structure is uniqueto van der Waals bonded solids in that significantstrain can be maintained even for very thick filmsgrown under nonequilibrium (i.e., quasi-epitaxial)

1890 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 99: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

conditions. Furthermore, the resulting thin-filmnanostructures have been used as “laboratories”useful in the investigation of fundamental optical andelectronic properties of the materials. These inves-tigations have led to a deeper understanding of thephysics of this enormous class of materials, while atthe same time have shown that precision growth ofthe thin films afforded by the UHV process of OMBDcan lead to significant modification of the materialsproperties which can be exploited in numerous deviceapplications. We have discussed several archetypedevices, and have shown that integrated organic/inorganic devices such as waveguides and photo-diodes, solar cells, organic light-emitting devices, andthin-film transistors based on OMBD-grown thinfilms are just now reaching performance and costlevels which make them extremely attractive for usein a large range of optoelectronic systems applica-tions.Given the relative youth of this technology, and the

exciting and rapid advances made using organicmolecular thin films, high-vacuum growth is poisedto make a significant contribution to our understand-ing and exploitation of these exciting materials wellinto the next century.

8.0. Acknowledgments

The author is indebted to numerous friends, col-laborators, and students who have contributed datato this review as well as to the author’s understand-ing of growth, phenomena, and applications of mo-lecular organic thin films. Foremost among thesecollaborators are Dr. M. L. Kaplan, Mr. P. H. Schmidt,and Dr. P. E. Burrows who have always freely sharedtheir ideas, talents, and enthusiasm of this subject,and who have contributed enormously to the under-standing and manipulation of organic molecularcrystals by the process of OMBD. I am also deeplygrateful for the contributions of many students.Among these, Dr. Franky So, Mr. Vladimir Bulovic,and Dr. Zilan Shen have made major advances in thedevelopment of OMBD science and technology. I alsothank Professor Mark Thompson for his many in-sights, as well as for synthesizing numerous materi-als used in our OLED structures. Many othercollaborators deserve acknowledgment, includingProfessor P. Eisenberger, Dr. P. Fenter, Mr. G. Gu,Dr. E. I. Haskal, Dr. Y. Hirose, Professor A. Kahn,and Dr. C. Kendrick. Numerous helpful discussionswith Professors S. Mukamel and Z. Soos are alsoacknowledged. The author thanks the Air ForceOffice of Scientific Research for their vision in sup-porting the effort in my laboratory long before therewas the slightest reason to believe that it wouldgenerate any practical outcome, as well as DARPA,the National Science Foundation, and UniversalDisplay Corporation. Finally, I express my deepestthanks to my wife Rosamund and my children fortheir nearly infinite patience and support of my work.

Figures that are cited with a reference number in thefigure caption have been reprinted with the permission ofthe publisher of the publication cited in the correspondingreference.

9.0. References(1) Forrest, S. R.; Burrows, P. E.; Haskal, E. I.; So, F. F. Phys. Rev.

B 1994, 49, 11309.(2) Hara, M.; Sasabe, H.; Yamada, A.; Garito, A. F. Jpn. J. Appl.

Phys. 1989, 28, L306.(3) Koma, A. Prog. Cryst. Growth Charact. 1995, 30, 129-152.(4) Tredgold, R. H. Order in Thin Organic Films; Cambridge

University Press: Cambridge, 1994.(5) Hong, H.; Tarabia, M.; Chayet, H.; Davidov, D.; Faraggi, E. Z.;

Avny, V.; Neumann, R.; Kirstein, S. J. Appl. Phys. 1996, 79,3082.

(6) Herman, M. A.; Sitter, H. Molecular Beam Epitaxy, 1st ed.;Springer-Verlag: Berlin, 1988.

(7) Optoelectronic Integration: Physics, Technology and Applica-tions; Wada, O., Ed.; Kluwer Academic Publishers: Boston, 1994.

(8) Fetterman, M.; Chao, C.-P.; Forrest, S. R. IEEE PhotonicsTechnol. Lett. 1995, 8, 69.

(9) Chao, C.-P.; Shiau, G.-J.; Forrest, S. R. IEEE Photonics Technol.Lett. 1994, 6, 1406.

(10) Sasaki, T.; Yamaguchi, M.; Kitamura, M. J. Cryst. Growth 1994,145, 846.

(11) Shiau, G.-J.; Chao, C.-P.; Burrows, P. E.; Forrest, S. R. Appl.Phys. Lett. 1994, 65, 892.

(12) Coldren, L. A.; Corzine, S. W. Diode Lasers and PhotonicIntegrated Circuits; John Wiley: New York, 1995.

(13) Miller, D. A. B.; Chemla, D. S.; Damen, T. C.; Gossard, A. C.;Wiegmann, W.; Wood, T. H.; Burrus, C. A. Phys. Rev. B 1985,32, 1043.

(14) Chemla, D. S.; Zyss, J. Nonlinear Optical Properties of OrganicMolecules and Crystals; Academic Press: Orlando, 1987.

(15) Mukamel, S. Principles of Nonlinear Optical Spectroscopy;Oxford University Press: New York, 1995.

(16) Organic Materials for Nonlinear Optics II; Hann, R. A., Bloor,D., Eds.; The Royal Soc. of Chemistry: Cambridge, 1991.

(17) Seto, J.; Asai, N.; Fujiwara, I.; Ishibashi, T.; Kamei, T.; Tamura,S. Thin Solid Films 1996, 273, 97.

(18) Maruyama, Y.; Hoshi, H.; Fang, S. L.; Kohama, K. Synth. Met.1995, 71, 1653.

(19) Simon, J.; Andre, J.-J. Molecular Semiconductors: Photoelec-trical Properties and Solar Cells; Springer-Verlag: Berlin, 1985.

(20) Wohrle, D.; Meissner, D. Adv. Mater. 1991, 3, 129.(21) Phthalocyanines: Properties and Applications; Leznoff, C. C.,

Lever, A. B. P., Eds.; VCH Publishers: New York, 1989.(22) Taylor, R. B.; Burrows, P. E.; Forrest, S. R. IEEE Photonics

Technol. Lett. 1997, 9, 365.(23) So, F. F.; Forrest, S. R.; Shi, Y. Q.; Steier, W. H. Appl. Phys.

Lett. 1990, 56, 674-676.(24) So, F. F.; Forrest, S. R. Phys. Rev. Lett. 1991, 66, 2649-2652.(25) Agranovich, V. M. Phys. Scr. 1993, 149, 699.(26) Agranovich, V. M.; LaRocca, G. C.; Bassani, F. Chem. Phys. Lett.

1995, 247, 355.(27) Lam, J. F.; Forrest, S. R.; Tangonan, G. L. Phys. Rev. Lett. 1991,

66, 1614.(28) Wang, N.; Jenkins, J. K.; Chernyak, V.; Mukamel, S. Phys. Rev.

B 1994, 49, 17079.(29) Zakhidov, A. A.; Yoshino, K. Synth. Met. 1994, 64, 155.(30) Nakada, H.; Tohma, T. 8th Int. Workshop on Electrolumines-

cence, Berlin, 1996; p 385.(31) Tang, C. W.; VanSlyke, S. A. Appl. Phys. Lett. 1987, 51, 913.(32) Burrows, P. E.; Forrest, S. R. Appl. Phys. Lett. 1994, 64, 2285.(33) Fang, S.; Kohama, K.; Hoshi, H.; Maruyama, Y. Synth. Met.

1994, 64, 167-170.(34) Arbour, C.; Armstrong, N. R.; Brina, R.; Collins, G.; Danziger,

J.; Dodelet, J.-P.; Lee, P.; Nebesny, K. W.; Pankow, J.; Waite,S. Mol. Cryst. Liq. Cryst. 1990, 183, 307.

(35) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H.; Venkatesan, T.;Lovinger, A. J. Appl. Phys. Lett. 1982, 41, 708.

(36) Lovinger, A. J.; Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H.;Venkatesan, T. J. Appl. Phys. 1984, 55, 476.

(37) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H. J. Appl. Phys. 1984,55, 1492.

(38) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H.; Feldmann, W. L.;Yanowski, E. Appl. Phys. Lett. 1982, 41, 90.

(39) Danziger, J.; Dodelet, J.-P.; Lee, P.; Nebesny, K. W.; Armstrong,N. R. Chem. Mater. 1991, 3, 821.

(40) Zimmermann, U.; Schnitzler, G.; Karl, N.; Umbach, E. ThinSolid Films 1989, 175, 85.

(41) Akers, K.; Aroca, R.; Hor, A.-M.; Loutfy, R. O. J. Phys. Chem.1987, 91, 2954.

(42) Ludwig, C.; Gompf, J.; Glatz, W.; Petersen, J.; Eisenmenger, W.;Mobus, M.; Karl, N. Z. Phys. B 1992, 86, 397.

(43) Hoshino, A.; Isoda, S.; Kurata, H.; Kobayashi, T. J. Appl. Phys.1994, 76, 4113.

(44) Vincett, P. S.; Popovic, Z. D.; McIntyre, L. Thin Solid Films 1981,82, 357.

(45) Debe, M. K.; Kam, K. K.; Liu, J. C.; Poirier, R. J. J. Vac. Sci.Technol. A 1988, 6, 1907.

(46) Mobus, M.; Karl, N.; Kobayashi, T. J. Cryst. Growth 1992, 116,495-504.

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1891

Page 100: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(47) So, F. F.; Forrest, S. R.; Shi, Y. Q.; Steier, W. H. Mat. Res. Soc.Ann. Mtg., 1990; p 71.

(48) So, F. F.; Leu, L. Y.; Forrest, S. R. SPIE Symp. on Growth ofSemiconductor Structures and High Tc Thin Films on Semicon-ductors, San Diego, 1990; p 95.

(49) Tanigaki, K.; Kuroshima, S.; Ebbesen, T. W.; Ichihashi, T. J.Cryst. Growth 1991, 114, 3.

(50) Yoshimura, T.; Tatsuura, S.; Sotoyama, W. Appl. Phys. Lett.1991, 59, 482.

(51) Tsang, W. T. In Lightwave Communications Technology; Tsang,W. T., Ed.; Academic Press: Orlando, 1985; Vol. 22A.

(52) Haskal, E. I.; So, F. F.; Burrows, P. E.; Forrest, S. R. Appl. Phys.Lett. 1992, 60, 3223-3225.

(53) Tada, H.; Kawaguchi, T.; Koma, A. Appl. Phys. Lett. 1992, 61,2021-2023.

(54) Hirose, Y.; Chen, W.; Haskal, E. I.; Forrest, S. R.; Kahn, A. J.Vac. Sci. Technol. B 1994, 12, 2616.

(55) Hirose, Y.; Forrest, S. R.; Kahn, A. Appl. Phys. Lett. 1995, 66,944-946.

(56) Zimmermann, U.; Karl, N. Surf. Sci. 1992, 268, 196.(57) Ludwig, C.; Gomph, B.; Petersen, J.; Strohmaier, R.; Eisen-

menger, W. Z. Phys. B 1994, 93, 365-373.(58) Kendrick, C.; Kahn, A.; Forrest, S. R. Appl. Surf. Sci. 1996, 104/

105, 586.(59) Gutmann, F.; Lyon, L. E. Rob’t. E. Krieger Publishing: Malabar,

FL, 1981.(60) Karl, N. Mol. Cryst. Liq. Cryst. 1989, 171, 157.(61) Warta, W.; Stehle, R.; Karl, N. Appl. Phys. A 1985, 36, 163.(62) Karl, N.; Marktanner, J.; Stehle, R.; Warta, W. Synth. Met. 1991,

41-43, 2473.(63) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H. InAnnu. Rev. Mater.

Sci.; Huggins, R. A., Ed.; Annual Reviews: Palo Alto, CA, 1987;Vol. 17.

(64) Weston, G. F. Butterworths: London, 1985.(65) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H. J. Appl. Phys. 1984,

56, 543.(66) Tanigaki, K.; Kuroshima, S.; Ebbesen, T.; Ichihashi, T. Mol.

Cryst. Liq. Cryst. Sci. Technol., Nonlinear Optics 1992, B2, 179-192.

(67) Matthews, J. W.; Blakeslee, A. E. J. Cryst. Growth 1974, 27,118.

(68) Kim, D. S.; Forrest, S. R.; Lange, M. J.; Olsen, G. H.; Cohen, M.J. IEEE Photonics Technol. Lett. 1994, 6, 235-237.

(69) Koma, A. Surf. Sci. 1992, 267, 29-33.(70) Burrows, P. E.; Zhang, Y.; Haskal, E. I.; Forrest, S. R.Appl. Phys.

Lett. 1992, 61, 2417.(71) Venables, J. A.; Spiller, G. D. T.; Hanbucken, M. Rep. Prog. Phys.

1984, 47, 399-459.(72) Zhang, Y.; Forrest, S. R. Phys. Rev. Lett. 1993, 71, 2765-2768.(73) Fenter, P.; Burrows, P. E.; Eisenberger, P.; Forrest, S. R. J. Cryst.

Growth 1995, 152, 65-72.(74) Fenter, P.; Eisenberger, P.; Burrows, P.; Forrest, S. R.; Liang,

K. S. Physica B 1996, 221, 145-151.(75) Novaco, A. D.; McTague, J. P. Phys. Rev. Lett. 1977, 38, 1286-

1289.(76) Jerome, B. Rep. Prog. Phys. 1991, 54, 391-451.(77) Shi, Y.; Cull, B.; Kumar, S. Phys. Rev. Lett. 1993, 71, 2773-

2776.(78) Forrest, S. R.; Burrows, P. E. Supramol. Sci. 1997, 4, 127.(79) Kitaigorodsky, A. I.Molecular Crystals and Molecules; Academic

Press: New York, 1973.(80) Ashcroft, N. W.; Mermin, N. D. Solid State Physics; Saunders

College: Philadelphia, 1976.(81) Yamamoto, H.; Tada, H.; Kawaguchi, T.; Koma, A. Appl. Phys.

Lett. 1994, 64, 2099-2101.(82) Tada, H.; Morioka, T.; Koma, A. J. Phys.: Condens. Matter 1994,

6, 1881-1892.(83) Tada, H.; Saiki, K.; Koma, A. Surf. Sci. 1992, 268, 387.(84) Hoshi, H.; Maruyama, Y. J. Appl. Phys. 1991, 69, 3046-3052.(85) Hoshi, H.; Fang, S.; Maruyama, Y. J. Appl. Phys. 1993, 73, 3111.(86) Dann, A. J.; Hoshi, H.; Maruyama, Y. J. Appl. Phys. 1990, 67,

1371.(87) Hoshi, H.; Maruyama, Y.; Masuda, H.; Inabe, T. J. Appl. Phys.

1990, 68, 1396.(88) Hoshi, H.; Dann, A. J.; Maruyama, Y. J. Appl. Phys. 1990, 67,

1845.(89) Hoshi, H.; Dann, A. J.; Maruyama, Y. J. Appl. Phys. 1990, 67,

6871.(90) Buchholz, J. C.; Somorjai, G. A. J. Chem. Phys. 1977, 66, 573.(91) Kawaguchi, T.; Tada, H.; Koma, A. J. Appl. Phys. 1994, 75,

1486-1492.(92) Schmicker, D.; Schmidt, S.; Skofronick, J. G.; Toennies, J. P.;

Vollmer, R. Phys. Rev. B 1991, 44, 10995-10997.(93) Krakow, W.; Rivera, N. M.; Roy, R. A.; Ruoff, R. S.; Cuomo, J. J.

J. Mater. Res. 1992, 7, 784.(94) Li, Y. Z.; Patrin, J. C.; Chander, M.; Weaver, J. H.; Chibante, L.

P. F.; Smalley, R. E. Science 1991, 252, 547.(95) Seidel, C.; Awater, C.; Liu, X. D.; Ellerbrake, R.; Fuchs, H. Surf.

Sci. 1997, 371, 123.

(96) Hoshino, A.; Isoda, S.; Kurata, H.; Kobayashi, T. J. Cryst. Growth1995, 146, 636.

(97) Tada, H.; Saiki, K.; Koma, A. Jpn. J. Appl. Phys. 1991, 30, L306.(98) Dietrich, T.; Schlesser, R.; Erler, B.; Kundig, A.; Sitar, Z.; Gunter,

P. J. Cryst. Growth 1997, 172, 473.(99) Forrest, S. R.; Zhang, Y. Phys. Rev. B 1994, 49, 11297-11308.(100) Burrows, P. E.; Forrest, S. R. Appl. Phys. Lett. 1993, 62, 3102-

3104.(101) Hirose, Y.; Forrest, S. R.; Kahn, A. Phys. Rev. B 1995, 52, 14040.(102) Zang, D. Y.; Shi, Y. Q.; So, F. F.; Forrest, S. R.; Steier, W. H.

Appl. Phys. Lett. 1991, 58, 562-564.(103) Komiyama, M.; Sakakibara, Y.; Hirai, H. Thin Solid Films 1987,

151, L109.(104) Hoshino, A.; Isoda, S.; Kurata, H.; Kobaysahi, T.; Yamashita,

Y. Jpn. J. Appl. Phys. 1995, 34, 3858.(105) Sakurai, M.; Tada, H.; Saiki, K.; Koma, A.; Funasaka, H.;

Kishimoto, Y. Chem. Phys. Lett. 1993, 208, 425-430.(106) Fartash, A. Appl. Phys. Lett. 1995, 67, 3901-3903.(107) Merwe, J. H. v. d. Philos. Mag. A 1982, 45, 127.(108) Reiss, H. J. Appl. Phys. 1968, 39, 5045.(109) Hashimoto, S.; Ogawa, T.; Asaka, N.; Isoda, S.; Kobayashi, T.

J. Cryst. Growth 1995, 146, 649.(110) Itoh, C.; Miyazaki, T.; Aizawa, K.; Aoki, H.; Okazaki, M. J. Phys.

C, Solid State Phys. 1988, 21, 4527-4544.(111) Fenter, P.; Schreiber, F.; Zhou, L.; Eisenberger, P. E.; Forrest,

S. R. Phys. Rev. B 1997, 56, 3046.(112) Abe, A.; Jernigan, R. L.; Flory, P. J. J. Am. Chem. Soc. 1966,

88, 631.(113) Scott, R. A.; A. Scheraga, H. J. Chem. Phys. 1965, 42, 2209.(114) Cox, J. D.; Pilcher, G. Thermochemistry of Organic and Orga-

nometallic Compounds; Academic Press: New York, 1970.(115) Taylor, R. B.; Shen, Z.; Forrest, S. R. Top. Mtg. on Organic Thin

Films for Photonics Applications Tech. Dig., 1995; p 38, PaperMB5.

(116) McTague, J. P.; Novaco, A. D. Phys. Rev. B 1979, 19, 5299-5306.

(117) Liu, D.-J.; Selinger, R. L. B.; Weeks, J. D. J. Chem. Phys. 1996,105, 4751.

(118) Dogo, S.; Germain, J.-P.; Maleysson, C.; Pauly, A. Thin SolidFilms 1992, 219, 244.

(119) Shimada, T.; Suzuki, A.; Sakurada, T.; Koma, A. Appl. Phys.Lett. 1996, 68, 2502.

(120) Hiller, A. C.; Ward, M. D. Phys. Rev. B 1996, 54, 14037.(121) Fecht, H. J. Philos. Mag. Lett. 1989, 60, 81.(122) Strohmaier, R.; Ludwig, C.; Petersen, J.; Gompf, B.; Eisen-

menger, W. Surf. Sci. 1996, 351, 292.(123) Kendrick, C. E. Princeton University, 1997.(124) Mobus, M.; Schreck, M.; Karl, N. Thin Solid Films 1989, 175,

89.(125) Seo, D. K.; Ren, J.; Wangbo, M. H. Surf. Sci. 1997, 370, 252.(126) Kotzian, M.; Rosch, N.; Schroeder, H.; Zerner, M. C. J. Am.

Chem. Soc. 1989, 111, 7687.(127) Anderson, W. P.; Edwards, W. D.; Zerner, M. C. Inorg. Chem.

1986, 25, 2728.(128) Anderson, W. P.; Cundari, T. R.; Drago, R. S.; Zerner, M. C.

Inorg. Chem. 1990, 29, 1.(129) Ohuchi, F. S.; Shimada, T.; Parkinson, B. A.; Ueno, K.; Koma,

A. J. Cryst. Growth 1991, 111, 1033-1037.(130) Sakurai, M.; Tada, H.; Saiki, K.; Koma, A. Jpn. J. Appl. Phys.

1991, 30, L1892-L1894.(131) Ueno, K.; Sakurai, M.; Koma, A. J. Cryst. Growth 1995, 150,

1180-1185.(132) Tanigaki, K.; Kuroshima, S.; Fujita, J.-I.; Ebbesen, T. W. Appl.

Phys. Lett. 1993, 63, 2351.(133) Nebesny, K. W.; Collins, G. E.; Lee, P. A.; Chau, L.-K.; Danziger,

J.; Osburn, E.; Armstrong, N. R. Chem. Mater. 1991, 3, 829.(134) Fritz, T.; Miskowiec, P. Mol. Cryst. Liq. Cryst. 1996, 283, 151.(135) Haskal, E. I.; Shen, Z.; Burrows, P. E.; Forrest, S. R. Phys. Rev.

B 1995, 51, 4449.(136) Dobson, P. J.; Joyce, B. A.; Neave, J. H.; Zhang, J. J. Cryst.

Growth 1987, 81, 1.(137) Cho, A. Y. J. Appl. Phys. 1971, 42, 2074.(138) Nonaka, T.; Mori, Y.; Nagai, N.; Nakagawa, Y.; Saeda, M.;

Takahagi, T.; Ishitani, A. Thin Solid Films 1994, 239, 214.(139) Nonaka, T.; Mori, Y.; Nagai, N.; Matsunobe, T.; Nakagawa, Y.;

Saeda, M.; Takahagi, T.; Ishitani, A. J. Appl. Phys. 1993, 73,2826.

(140) Fuchigami, H.; Tanimura, S.; Uehara, Y.; Kurata, T.; Tsunoda,S. Jpn. J. Appl. Phys. 1995, 34, 3852.

(141) Collins, G. E.; Williams, V. S.; Chau, L.-K.; Nebesny, K. W.;England, C.; Lee, P. A.; Lowe, T.; Fernando, Q.; Armstrong, N.R. Synth. Met. 1993, 54, 351.

(142) Imanishi, Y.; Hattori, S.; Kakuta, A.; Numata, S. Phys. Rev. Lett.1993, 71, 2098.

(143) Nonaka, T.; Date, T.; Tomita, S.; Nagai, N.; Nishimura, M.;Murata, Y.; Ishitani, A. Thin Solid Films 1994, 237, 87.

(144) Takada, J.; Awaji, H.; Koshioka, M.; Nevin, W. A. J. Appl. Phys.1994, 75, 4055.

(145) Tokito, S.; Sakata, J.; Taga, Y. J. Appl. Phys. 1995, 77, 1985-1989.

1892 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 101: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(146) Ohmori, Y.; Fujii, A.; Uchida, M.; Morishima, C.; Yoshino, K.Appl. Phys. Lett. 1993, 62, 3250.

(147) Ohmori, Y.; Fujii, A.; Uchida, M.; Morishima, C.; Yoshino, K.Appl. Phys. Lett. 1993, 63, 1871.

(148) Fujii, A.; Yoshida, M.; Ohmori, Y.; Yoshino, K. Jpn. J. Appl. Phys.1995, 34, L621.

(149) Fujii, A.; Yoshida, M.; Ohmori, Y.; Yoshino, K. Jpn. J. Appl. Phys.1995, 34, L499.

(150) Hong, H.; Tarabia, M.; Chayet, H.; Davidov, D.; Faraggi, E. Z.;Avny, V.; Neumann, R.; Kirstein, S. J. Appl. Phys. 1996, 79,3082.

(151) Armstrong, N. R.; Nebesny, K. W.; Collins, G. E.; Chau, L.-K.;Lee, P. A.; England, C.; Diehl, D.; Douskey, M.; Parkinson, B.A. Thin Solid Films 1992, 216, 90.

(152) Nanai, N.; Yudasaka, M.; Ohki, Y.; Yoshimura, S. Thin SolidFilms 1995, 265, 1.

(153) Born, L.; Heywang, G. Z. Kristallogr. 1990, 190, 153.(154) Burrows, P. E.; Hara, M.; Sasabe, H.Mol. Cryst. Liq. Cryst. Sci.

Technol. B, Nonlinear Opt. 1992, 2, 193.(155) Ludwig, C.; Strohmaier, R.; Petersen, J.; Gompf, B.; Eisen-

menger, W. J. Vac. Sci. Technol. B 1994, 12, 1963.(156) Nagao, S.; Takashima, M.; Inoue, Y.; Katoh, M.; Gotoh, H. J.

Cryst. Growth 1991, 111, 521.(157) Lyons, M. H. J. Cryst. Growth 1989, 96, 339.(158) Shiau, G.-J.; Chao, C.-P.; Burrows, P. E.; Forrest, S. R. J. Appl.

Phys. 1995, 77, 201.(159) Tokito, S.; Sakata, J.; Taga, Y. Appl. Phys. Lett. 1994, 64, 1353.(160) Yudson, V. I.; Reineker, P.; Agranovich, V. M. Phys. Rev. B 1995,

52, R5543.(161) Vlieg, E.; Gon, A. W. D. v. d.; Veen, J. F. v. d.; Macdonald, J. E.;

Norris, C. Phys. Rev. Lett. 1988, 61, 2241.(162) Mobus, M.; Karl, N. Thin Solid Films 1992, 215, 213-217.(163) Haskal, E. I.; Zhang, Y.; Burrows, P. E.; Forrest, S. R. Chem.

Phys. Lett. 1994, 219, 325.(164) Li, D.; Nakamura, Y.; Otsuka, N.; Qiu, J.; Kobayashi, M.;

Gunshor, R. L. J. Vac. Sci. Technol. B 1991, 9, 2167.(165) Li, D.; Pashley, M. D. Phys. Rev. B 1994, 49, 13643.(166) Pashley, M. D.; Li, D. J. Vac. Sci. Technol. A 1994, 12, 1848.(167) Hirose, Y.; Chen, W.; Haskal, E. I.; Forrest, S. R.; Kahn, A. Appl.

Phys. Lett. 1994, 64, 3482.(168) Chadi, D. J. J. Vac. Sci. Technol. A 1987, 5, 834.(169) Larsen, P. K.; Veen, J. F. v. d.; Mazur, A.; Pollmann, J.; Neave,

J. H.; Joyce, B. A. Phys. Rev. B 1982, 26, 3222.(170) So, F. F.; Forrest, S. R. J. Appl. Phys. 1988, 63, 442.(171) Forrest, S. R.; So, F. F. J. Appl. Phys. 1988, 64, 399.(172) Bulovic, V.; Forrest, S. R. Chem. Phys. 1996, 210, 13.(173) Bulovic, V.; Forrest, S. R. Chem. Phys. Lett. 1995, 238, 88.(174) Forrest, S. R.; Yoon, W. Y.; Leu, L. Y.; So, F. F. J. Appl. Phys.

1989, 66, 5908.(175) Zang, D. Y.; So, F. F.; Forrest, S. R. Appl. Phys. Lett. 1991, 59,

823-825.(176) Shen, Z.; Forrest, S. R. Phys. Rev. B 1997, 55, 10578.(177) Shen, Z.; Burrows, P. E.; Forrest, S. R.; Ziari, M.; Steier, W. H.

Chem. Phys. Lett. 1995, 236, 129.(178) Sakurai, M.; Koma, A. Trans. Mater. Res. Soc. Jpn. 1994, 14B,

1145-1148.(179) Nagayama, K.; Mitsumoto, R.; Araki, T.; Ouchi, Y.; Seki, K.

Physica B 1995, 208-209, 419.(180) Gimzewski, J. K.; Modesti, S.; Gerber, C.; Schlittler, R. R. Chem.

Phys. Lett. 1993, 213, 401-406.(181) Altman, E. I.; Colton, R. J. Surf. Sci. 1992, 279, 49-67.(182) Viswanathan, R.; Zasadzinski, J. A.; Schwartz, D. K. Science

1993, 261, 449-452.(183) Akimichi, H.; Inoshita, T.; Hotta, S.; Noge, H.; Sakaki, H. Appl.

Phys. Lett. 1993, 63, 3158.(184) Yoshimura, T.; Tatsuura, S.; Sotoyama, W. Thin Solid Films

1992, 207, 9.(185) Ahonen, M.; Pessa, M.; Suntola, T. Thin Solid Films 1980, 65,

301.(186) Bedair, S. M.; Tischler, M. A.; Katsuyama, T.; El-Masry, N. A.

Appl. Phys. Lett. 1985, 47, 51.(187) Ide, Y.; McDermott, B. T.; Hashemi, M.; Bedair, S. M.; Goodhue,

W. D. Appl. Phys. Lett. 1988, 53, 2314.(188) Isshiki, H.; Aoyagi, Y.; Sugano, T.; Iwai, S.; Meguro, T. Appl.

Phys. Lett. 1993, 63, 1528.(189) An, N. B. Phys. Lett. A 1996, 214, 99.(190) Takada, J. Jpn. J. Appl. Phys. 1995, 34, 3864.(191) Takada, J.; Awaji, H.; Koshioka, M.; Nakajima, A.; Nevin, W.

A. Appl. Phys. Lett. 1992, 61, 2184.(192) Hasegawa, S.; Mori, T.; Imaeda, K.; Tanaka, S.; Yamashita, Y.;

Inokuchi, H.; Fujimoto, H.; Seki, K.; Ueno, N. J. Chem. Phys.1994, 100, 6969.

(193) Inokuchi, H.; Imaeda, K. Acta Phys. Pol. A 1995, 88, 1161.(194) Imeada, K.; Yamashita, Y.; Li, Y.; Mori, T.; Inokuchi, H.; Sano,

M. J. Mater. Chem. 1992, 2, 115.(195) Silinsh, E. A. Organic Molecular Crystals: Their Electronic

States; Springer-Verlag: Berlin, 1980.(196) Pope, M.; Swenberg, C. E. Electronic Processes in Organic

Crystals; Clarendon Press: Oxford, 1982.

(197) Bulovic, V.; Burrows, P. E.; Forrest, S. R.; Cronin, J. A.;Thompson, M. E. Chem. Phys. 1996, 210, 1.

(198) Kelly, M. K.; Etchegoin, P.; Fuchs, D.; Kratschmer, W.; Fos-tiropoulos, K. Phys. Rev. B 1992, 46, 4963.

(199) Sebastian, L.; Weiser, G.; Bassler, H. Chem. Phys. 1981, 61, 125.(200) Tanaka, J. Bull. Chem. Soc. Jpn. 1963, 36, 1237.(201) Schreiber, M.; Toyozawa, Y. J. Phys. Soc. Jpn. 1982, 51, 1544.(202) Mataga, N.; Torihashi, Y.; Ota, Y. Chem. Phys. Lett. 1967, 1,

385.(203) So, F. F.; Forrest, S. R. Mol. Cryst. Liq. Cryst. Sci. Technol. B:

Nonlinear Optics 1992, 2, 205-222.(204) Anderson, M. L.; Williams, V. S.; Schuerlein, T. J.; Collins, G.

E.; England, C. D.; Chau, L.-K.; Lee, P. A.; Nebesny, K. W.;Armstrong, N. R. Surf. Sci. 1994, 307-309, 551.

(205) Weisbuch, C.; Dingle, R.; Gossard, A. C.; Wiegmann, W. SolidState Commun. 1981, 38, 709.

(206) Broude, V. L.; Rashba, E. I.; Sheka, E. F. Spectroscopy ofMolecular Excitons; Springer-Verlag: Berlin, 1985.

(207) Sugano, S.; Tanabe, Y.; Kamimura, H. Multiplets of TransitionMetal Ions in Crystals; Academic: New York, 1970.

(208) Pikus, G. E.; Bir, G. L. Symmetry and Deformation Effects inSemiconductors; Nauka: Moscow, 1972 (English Translation,Wiley: New York, 1974).

(209) Cho, K. Excitons; Springer-Verlag: Berlin, 1979.(210) Silinsh, E. A.; Capek, V. Organic Molecular Crystals; American

Institute of Physics: New York, 1994.(211) Ueta, M.; Kanazaki, H.; Kobayashi, K.; Toyozawa, Y. Excitonic

Processes in Solids; Springer-Verlag: New York, 1986.(212) Sebastian, L.; Weiser, G.; Bassler, H. Chem. Phys. 1981, 61, 125.(213) Bounds, P. J.; Petelenz, P.; Siebrand, W. Chem. Phys. 1981, 63,

303.(214) Hasegawa, S.; Tanaka, S.; Yamashita, Y.; Inokuchi, H.; Fujimoto,

H.; Kamiya, K.; Seki, K.; Ueno, N. Phys. Rev. B 1993, 48, 2596.(215) Wu, C. I.; Hirose, Y.; Sirringhaus, H.; Kahn, A. Chem. Phys. Lett.

1997.(216) Hirose, Y.; Aristov, V.; Soukiassian, P.; Bulovic, V.; Forrest, S.

R.; Kahn, A. Phys. Rev. B 1996, 54, 13748.(217) So, F. F.; Forrest, S. R. Appl. Phys. Lett. 1988, 52, 1341.(218) Dubovsky, O.; Mukamel, S. J. Chem. Phys. 1992, 15, 417.(219) Huang, H. X.; Mukamel, S. Chem. Phys. Lett. 1992, 192, 417.(220) Mukamel, S.; Takahashi, A.; Huang, H. X.; Chen, G. Science

1994, 266, 250.(221) Wang, N.; Mukamel, S. Chem. Phys. Lett. 1994, 231, 373.(222) Silbey, R.; Jortner, J.; Rice, S. A.; Vala, M. T. J. Chem. Phys.

1965, 43, 2925.(223) Silbey, R.; Jortner, J.; Rice, S. A.; Vala, M. T. J. Chem. Phys.

1965, 42, 733.(224) Ostrick, J. R.; Dodabalapur, A.; Torsi, L.; Lovinger, A. J.; Kwock,

E. W.; Miller, T. M.; Galvin, M.; Berggren, M.; Katz, H. E. J.Appl. Phys. 1997, 81, 6804.

(225) Jurgis, A.; Silinsh, E. A. Phys. Status Solidi B 1972, 53, 735.(226) Bounds, P. J.; Siebrand, W. Chem. Phys. Lett. 1980, 75, 414.(227) Mott, N. F.; Gurney, R. W. Electronic Processes in Ionic Crystals;

Clarendon: Oxford, 1940.(228) Toyozawa, Y. Prog. Theor. Phys. 1954, 12, 421.(229) Bounds, P. J.; Munn, R. W. Chem. Phys. 1979, 44, 103.(230) Jackson, J. D. Classical Electrodynamics, 2nd ed.; Wiley: New

York, 1975.(231) Shinozuka, Y.; Matsuura, M. Phys. Rev. B 1983, 28, 4878.(232) Ruff, M.; Streb, D.; Dankowski, S. U.; Tautz, S.; Kiesel, P.;

Knupfer, B.; Kneissl, M.; Linder, N.; Dohler, G. H.; Keil, U. D.Appl. Phys. Lett. 1996, 68, 2968.

(233) Sebastian, L.; Weiser, G. Chem. Phys. 1981, 62, 447.(234) Meyling, J. H.; Wiersma, D. A. Chem. Phys. Lett. 1973, 20, 383.(235) Sebastian, L.; Weiser, G. Chem. Phys. 1981, 62, 447.(236) Schiff, L. I. QuantumMechanics; McGraw-Hill: New York, 1968.(237) Sebastian, L.; Weiser, G. Phys. Rev. Lett. 1981, 46, 1156.(238) Boer, K. W. Survey of Semiconductor Physics: Electrons and

Other Particles in Bulk Semiconductors; Van Nostrand Rein-hold: New York, 1990.

(239) Laporte, P.; Subtil, J. L. Phys. Rev. Lett. 1980, 45, 2138.(240) Jayaraman, A.; Kaplan, M. L.; Schmidt, P. H. J. Chem. Phys.

1985, 82, 1682.(241) Matsui, A.; Mizuno, K. Chem. Phys. 1987, 113, 111.(242) Goebel, E. O.; Jung, H.; Kuhl, J.; Ploog, K. Phys. Rev. Lett. 1983,

51, 1588.(243) Bebb, H. B.; Williams, E. W. In Semiconductors and Semimetals;

Willardson, R. K., Beer, A. C., Eds.; Academic Press: New York,1978; Vol. 8.

(244) Haarer, D. Chem. Phys. Lett. 1974, 27, 91.(245) Munn, R. W.; Siebrand, W. J. Chem. Phys. 1970, 52, 47.(246) Hong, H.; Davidov, D.; Avny, Y.; Chayet, H.; Faraggi, E. Z.;

Neumann, R. Adv. Mater. 1995, 7, 846.(247) Munn, R. W. J. Chem. Phys. 1994, 99, 1404.(248) Agranovich, V. M.; Atanasov, R. D.; Bassani, G. F. Chem. Phys.

Lett. 1992, 199, 621.(249) An, N. B.; Hanamura, E. J. Phys.: Condens. Matter 1996, 8,

2273.(250) Shen, Z.; Forrest, S. R. Phys. Rev. B 1993, 48, 17584.

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1893

Page 102: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(251) Yudson, V. I.; Neidlinger, T.; Reineker, P. Phys. Lett. A 1995,204, 313.

(252) Hong, H.; Davidov, D.; Chayet, H.; Faraggi, E. Z.; Tarabia, M.;Avny, Y.; Neumann, R.; Kirstein, S. Supramol. Sci. 1997, 4, 67.

(253) Chen, X. L.; Jenekhe, S. A. Appl. Phys. Lett. 1997, 70, 487.(254) Tokura, Y.; Koda, T.; Iyechika, Y.; Kuroda, H. Chem. Phys. Lett.

1983, 102, 174.(255) Yoshida, H.; Tokura, Y.; Koda, T. Chem. Phys. 1986, 109, 375.(256) Gu, G.; Bulovic, V.; Burrows, P. E.; Forrest, S. R.; Thompson,

M. E. Appl. Phys. Lett. 1996, 68, 2606-2608.(257) Burrows, P. E.; Sapochak, L. S.; McCarty, D. M.; Forrest, S. R.;

Thompson, M. E. Appl. Phys. Lett. 1994, 64, 2718.(258) Burrows, P. E.; Shen, Z.; Bulovic, V.; McCarty, D. M.; Forrest,

S. R.; Cronin, J. A.; Thompson, M. E. J. Appl. Phys. 1996, 79,7991.

(259) Petelenz, P.; Slawik, M.; Pac, B. Synth. Met. 1994, 64, 335.(260) Fischer, C. M.; Burghard, M.; Roth, S. Synth. Met. 1995, 71,

1975.(261) Fujii, A.; Yoshida, M.; Ohmori, Y.; Yoshino, K. Thin Solid Films

1996, 273, 199.(262) Ghosh, A. K.; Feng, T. J. Appl. Phys. 1978, 49, 5982.(263) Karl, N.; Baurer, A.; Holzapfel, J.; Marktanner, J.; Mobus, M.;

Stolzle, F. Mol. Cryst. Liquid Cryst. 1994, 252, 243.(264) Tang, C. W.; VanSlyke, S. A.; Chen, C. H. J. Appl. Phys. 1989,

65, 3610.(265) Torsi, L.; Dodabalapur, A.; Rothberg, L. J.; Fung, A. W. P.; Katz,

H. E. Science 1996, 272, 1462.(266) Dodabalapur, A.; Katz, H. E.; Torsi, L.; Haddon, R. C. Science

1995, 269, 1560.(267) Dodabalapur, A.; Torsi, L.; Katz, H. E. Science 1995, 268, 270.(268) Garnier, F.; Horowitz, G.; Fichou, D.; Yassar, A. Supramol. Sci.

1997, 4, 155.(269) Lin, Y.-Y.; Gundlach, D. J.; Jackson, T. N. Mat. Res. Soc. Spring

Mtg., San Francisco, CA, 1997; p 143.(270) Garnier, F.; Horowitz, G.; Peng, X.; Fichou, D. Adv. Mater. 1990,

2, 1990.(271) Tang, C. W. Appl. Phys. Lett. 1986, 48, 183.(272) Fang, S.; Hoshi, H.; Kohama, K.; Maruyama, Y. J. Phys. Chem.

1996, 100, 4104.(273) Fang, S.; Kohama, K.; Hoshi, H.; Maruyama, Y. Chem. Phys.

Lett. 1995, 234, 343.(274) Fang, S.; Kohama, K.; Hoshi, H.; Maruyama, Y. Jpn. J. Appl.

Phys. 1993, 32, L1418.(275) Hoshi, H.; Kohama, K.; Fang, S.; Maruyama, Y. Appl. Phys. Lett.

1993, 62, 3080.(276) Leegwater, J. A.; Mukamel, S. Phys. Rev. A 1992, 46, 452.(277) Dubovsky, O.; Mukamel, S. J. Chem. Phys. 1991, 95, 7828.(278) Samoc, M.; Prasad, P. N. J. Chem. Phys. 1989, 91, 6643.(279) Yamamoto, H.; Katogi, S.; Watanabe, T.; Sato, H.; Miyata, S.;

Hosomi, T. Appl. Phys. Lett. 1992, 60, 935.(280) Zang, D. Y.; Forrest, S. R. Appl. Phys. Lett. 1992, 60, 189-191.(281) Zang, D. Y.; Forrest, S. R. IEEE Photonics Technol. Lett. 1992,

4, 365.(282) So, F. F.; Forrest, S. R. IEEE Trans. Electron. Dev. 1989, 36,

66.(283) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H.; J. M. Parsey, J. J.

Appl. Phys. 1985, 58, 867.(284) Schmidt, P. H.; Forrest, S. R.; Kaplan, M. L. J. Electrochem.

Soc. 1986, 133, 769.(285) Forrest, S. R.; Schmidt, P. H.; Wilson, R. B.; Kaplan, M. L. J.

Vac. Sci. Technol. B 1986, 4, 37.(286) Chang, S. J.; Leu, L. Y.; Forrest, S. R.; Jones, C. E. Appl. Phys.

Lett. 1989, 54, 1040.(287) Al-Mohamad, A.; Smith, C. W.; Al-Saffar, I. S.; Salfin, M. A. Thin

Solid Films 1990, 175, 181.(288) Al-Mohamad, A.; Soukieh, M. Thin Solid Films 1995, 271, 132.(289) Forrest, S. R.; Schmidt, P. H.; Wilson, R. B.; Kaplan, M. L. Appl.

Phys. Lett. 1984, 45, 1199.(290) Forrest, S. R.; Kaplan, M. L.; Schmidt, P.; Gates, J. V. J. Appl.

Phys. 1985, 57, 2982.(291) Forrest, S. R.; Schmidt, P. H. J. Appl. Phys. 1986, 59, 513.(292) Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H. J. Appl. Phys. 1986,

60, 2406.(293) Antohe, S.; Tomozeiu, N.; Gogonea, S. Phys. Status Solidi 1991,

125, 397.(294) Prabhakar, V.; Forrest, S. R.; Lorenzo, J. P.; Vaccaro, K. IEEE

Photonics Technol. Lett. 1990, 2, 724.(295) Adachi, C.; Nagai, K.; Tamoto, N. Appl. Phys. Lett. 1995, 66,

2679.(296) Yap, D.; Burrows, P. E.; Forrest, S. R. Top. Mtg. Organic Thin

Films for Photonics Applications Tech. Dig. 1995, Paper WB2,p 302.

(297) Cheng, C.-L.; Forrest, S. R.; Kaplan, M. L.; Schmidt, P. H.; Tell,B. Appl. Phys. Lett. 1985, 47, 1217.

(298) Johnson, J. G.; Forrest, S. R.; Zeisse, C. R.; Nguyen, R. Appl.Phys. Lett. 1988, 52, 495.

(299) Chamberlain, G. A.; Clooney, P. J. Chem. Phys. Lett. 1979, 66,88.

(300) Forrest, S. R.; Yoon, W. Y.; Leu, L. Y.; So, F. F. J. Appl. Phys.1989, 66, 5908.

(301) Gregg, B. A. Appl. Phys. Lett. 1995, 67, 1271.(302) Hiramoto, M.; Suezaki, M.; Yokoyama, M. Chem. Lett. 1990, 327.(303) Hiramoto, M.; Kishigami, Y.; Yokoyama, M. Chem. Lett. 1990,

119.(304) Hiramoto, M.; Fukusumi, H.; Yokoyama, M. Appl. Phys. Lett.

1992, 61, 2580.(305) Karl, N.; Bauer, A.; Holzapfel, J.; Marktanner, J.; Mobus, M.;

Stozle, F. Mol. Cryst. Liq. Cryst. 1994, 252, 243.(306) Siebentritt, S.; Gunster, S.; Meissner, D. Synth. Met. 1991, 41-

43, 1173.(307) Wohrle, D.; Kreienhoop, L.; Schnurpfeil, G.; Elbe, J.; Tennigkeit,

B.; Hiller, S.; Schlettwein, D. J. Mater. Chem. 1995, 5, 1819.(308) Yamashita, K.; Harima, Y.; Matsubayashi, T. J. Phys. Chem.

1989, 93, 5311.(309) Panayotatos, P.; Bird, G.; Sauers, R.; Piechowski, A.; Husain,

S. Sol. Cells 1987, 21, 301.(310) Gunster, S.; Siebentritt, S.; Elbe, J.; Kreienhoop, L.; Tennigkeit,

B.; Wohrle, D.; Memming, R.; Meissner, D.Mol. Cryst. Liq. Cryst.1992, 218, 117.

(311) Halls, J. J. M.; Pichler, K.; Friend, R. H.; Moratti, S. C.; Holmes,A. B. Appl. Phys. Lett. 1996, 68, 3120.

(312) Yonehara, H.; Pac, C. Thin Solid Films 1996, 278, 108.(313) Sze, S. M. Physics of Semiconductor Devices, 2 ed.; John Wiley:

New York, 1981.(314) Heterojunction Band Discontinuities: Physics and Device Ap-

plications; Capasso, F., Margaritondo, G., Eds.; North-Holland:Amsterdam, 1987.

(315) Kao, K. C.; Hwang, W. Electrical Transport in Solids; Perga-mon: Oxford, 1981.

(316) Chaiken, R. F.; Kearns, D. R. J. Chem. Phys. 1966, 45, 3966.(317) Garbuzov, D. Z.; Forrest, S. R.; Tsekoun, A. G.; Burrows, P. E.;

Bulovic, V.; Thompson, M. E. J. Appl. Phys. 1996, 80, 4644.(318) Blouke, M. M.; Cowens, M. W.; Hall, J. E.; Westphal, J. A.;

Christensen, A. B. Appl. Opt. 1980, 19, 3318.(319) Goetzberger, A.; Greubel, W. Appl. Phys. 1977, 14, 123.(320) Viehmann, W. Conf. Meas. Opt. Rad. 1979; p 90.(321) Sims, G.; Griffin, F.; Lesser, M. Opt. Sensors and Electron.

Photog. 1989; p 31.(322) Garbuzov, D. Z.; Bulovic, V.; Burrows, P. E.; Forrest, S. R. Chem.

Phys. Lett. 1996, 249, 433.(323) Dresner, J. RCA Rev. 1969, 30, 322.(324) Helfrich, W.; Schneider, W. G. Phys. Rev. Lett. 1965, 14, 229.(325) Tang, C. W. Inf. Dis. 1996, 16.(326) Wu, C. C.; Chun, J. K. M.; Burrows, P. E.; Sturm, J. C.;

Thompson, M. E.; Forrest, S. R.; Register, R. A. Appl. Phys. Lett.1995, 66, 653.

(327) Era, M.; Morimoto, S.; Tsutsui, T.; Saito, S. Appl. Phys. Lett.1994, 65, 676.

(328) Forrest, S. R.; Burrows, P. E.; Thompson, M. E. Laser FocusWorld 1995, 99.

(329) Sheats, J. R.; Antoniadis, H.; Hueschen, M. R.; Leonard, W.;Miller, J. N.; Moon, R. L. Science 1996, 273, 884.

(330) Rothberg, L. J.; Lovinger, A. J. J. Mater. Res. 1996, 11, 3174.(331) Greenham, N. C.; Friend, R. H. Solid State Phys. 1995, 49, 1.(332) Brown, A. R.; Bradley, D. D. C.; Burroughes, J. H.; Friend, R.

H.; Greenham, N. C.; Burns, P. L.; Holmes, A. B.; Kraft, A. Appl.Phys. Lett. 1992, 61, 2793.

(333) Burroughes, J. H.; Bradley, D. D. C.; Brown, A. R.; Marks, R.N.; Mackay, K.; Friend, R. H.; Burns, P. L.; Holmes, A. B.Nature1990, 347, 539.

(334) Sato, Y.; Kanai, H. Mol. Cryst. Liq. Cryst. 1994, 253, 143.(335) Do, L. M.; Han, E. M.; Niidome, Y.; Fujihira, M.; Kanno, T.;

Yoshida, S.; Maeda, A.; Iksushima, A. J. J. Appl. Phys. 1994,76, 5118.

(336) McElvain, J.; Antoniadis, H.; Hueschen, M. R.; Miller, J. N.;Roitman, D. M.; Sheats, J. R.; Moon, R. L. J. Appl. Phys. 1996,80, 6002.

(337) Burrows, P. E.; Bulovic, V.; Forrest, S. R.; Sapochak, L. S.;McCarty, D. M.; Thompson, M. E. Appl. Phys. Lett. 1994, 65,2922.

(338) VanSlyke, S. A.; Chen, C. H.; Tang, C. W. Appl. Phys. Lett. 1996,69, 2160.

(339) Shi, J.; Tang, C. W. Appl. Phys. Lett. 1997, 70, 1665.(340) Han, E.-M.; Do, L.-M.; Nidome, Y.; Fujihira, M. Chem. Lett. 1994,

969.(341) Fenter, P.; Schreiber, F.; Bulovic, V.; Forrest, S. R. Chem. Phys.

Lett. 1997, in press.(342) Orita, K.; Morimura, T.; Horiuchi, T.; Matsushige, K. Int. Conf.

on Electrolum. of Molec. Mat. and Related Phenom., Kitakyushu,Japan, 1997; p 106.

(343) Tidswell, I. M.; Ocko, B. M.; Pershan, P. S.; Wasserman, S. R.;Whitesides, G. M.; Axe, J. D. Phys. Rev. B 1990, 41, 1111.

(344) Tokito, S.; Tanaka, H.; Okada, A.; Taga, Y. Appl. Phys. Lett.1996, 69, 878.

(345) Papadimitrakopoulos, F.; Zhang, X.-M.; Thomsen, D. L., III;Higginson, K. A. Chem. Mater. 1996, 8, 1363.

(346) Shen, Z.; Burrows, P. E.; Bulovic, V.; McCarty, D. M.; Thompson,M. E.; Forrest, S. R. Jpn. J. Appl. Phys. Lett. 1996, 35, L401.

(347) Garbuzov, D. Z.; Bulovic, V.; Burrows, P. E.; Forrest, S. R. Chem.Phys. Lett. 1996, 249, 433.

1894 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest

Page 103: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(348) Adachi, C.; Tsutsui, T.; Saito, S. Optoelectron.sDev. Technol.1991, 6, 25.

(349) Kalinowski, J.; Giro, G.; Marco, P. D.; Camaioni, N.; Fattori, V.Chem. Phys. Lett. 1997, 265, 607.

(350) Lin, L.-B.; Jenekhe, S. A.; Young, R. H.; Borsenberger, P. M.Appl. Phys. Lett. 1997, 70, 2052.

(351) Hung, L. S.; Tang, C. W.; Mason, M. G. Appl. Phys. Lett. 1997,70, 152.

(352) Wu, C.-C.; Chun, J.; Burrows, P.; Forrest, S. R.; Register, R. A.;Sturm, J. C.; Thompson, M. E. Polym. Prepr. 1994, 35, 101.

(353) Suzuki, H.; Hoshino, S. J. Appl. Phys. 1996, 79, 858.(354) Suzuki, H.; Hoshino, S. J. Appl. Phys. 1996, 79, 8816.(355) Blom, P. W. M.; deJong, M. J. M.; Vleggar, J. J. M. Appl. Phys.

Lett. 1996, 68, 3308.(356) Lampert, M. A.; Mark, P. Current Injection In Solids; Aca-

demic: New York, 1970.(357) Friedman, L. Phys. Rev. A 1965, 140, 1649.(358) Hosokawa, C.; Tokailin, H.; Higashi, H.; Kusumoto, T. Appl.

Phys. Lett. 1992, 60, 1220.(359) Shockley, W.; Read, W. T. Phys. Rev. 1952, 87, 835.(360) Forrest, S. R.; Shen, Z.; Burrows, P. E.; Bulovic, V.; Thompson,

M. E. Inorganic and Organic Electroluminescence/EL ’96, Berlin,1996; p 83.

(361) Burrows, P. E.; Forrest, S. R.; Sibley, S. P.; Thompson, M. E.Appl. Phys. Lett. 1996, 69, 2959.

(362) Burrows, P. E.; Forrest, S. R.; Sibley, S. M.; Thompson, M. E.Mat. Res. Soc. Ann. Mtg., Boston, 1996; p D3.10.

(363) Matsumura, M.; Akai, T.; Saito, M.; Kimura, T. J. Appl. Phys.1996, 79, 264.

(364) Matsumura, M.; Jinde, Y.; Akai, T.; Kimura, T. Jpn. J. Appl.Phys. 1996, 35, 5735.

(365) Parker, I. D. J. Appl. Phys. 1994, 75, 1656.(366) Joswick, M. D.; Campbell, I. H.; Barashkov, N. N.; Ferraris, J.

P. J. Appl. Phys. 1996, 80, 2883.(367) Bulovic, V.; Tian, P.; Burrows, P. E.; Gokhale, M. R.; Forrest, S.

R.; Thompson, M. E. Appl. Phys. Lett. 1997, 70, 294.(368) Sugiyama, K.; Yoshimura, D.; Ito, E.; Miyazaki, T.; Hamatani,

Y.; Kawamoto, I.; Ouchi, Y.; Seki, K.; Ishii, H. Mol. Cryst. Liq.Cryst. 1996, 285-286, 239.

(369) Hirose, Y.; Kahn, A.; Aristov, V.; Soukiassian, P.; Bulovic, V.;Forrest, S. R. Phys. Rev. B 1996, 54, 13748.

(370) Narioka, S.; Ishii, H.; Yoshimura, D.; Sei, M.; Ouchi, Y.; Seki,K.; Hasegawa, S.; Miyazaki, T.; Harima, Y.; Yamashita, K. Appl.Phys. Lett. 1995, 67, 1899.

(371) Probst, M.; Haight, R. Appl. Phys. Lett. 1997, 70, 1420.(372) Mori, Y. In Organic Electroluminscent Materials and Devices;

Miyata, S., Nalwa, H. S., Eds.; Gordon & Breach: Amsterdam,1997.

(373) Burrows, P. E.; Gu, G.; Bulovic, V.; Forrest, S. R.; Thompson,M. E. IEEE Trans. Electron. Dev. Lett. 1997, 44.

(374) Wu, C. C.; Sturm, J. C.; Register, R. A.; Thompson, M. E. Appl.Phys. Lett. 1996, 69, 3117.

(375) Strukelj, M.; Jordan, R. H.; Dodabalapur, A. J. Am. Chem. Soc.1996, 118, 1213.

(376) Kido, J.; Kimura, M.; Nagai, K. Science 1995, 267, 1332.(377) Kido, J.; Ikeda, W.; Kimura, M.; Nagai, K. Jpn. J. Appl. Phys.

1996, 35, L394.(378) Hamada, Y.; Adachi, C.; Tsutsui, T.; Saito, S. Jpn. J. Appl. Phys.

1992, 31, 1812.(379) Matsura, M.; Tokailin, H.; Eida, M.; Hosokawa, C.; Hironaka,

Y.; Kusumoto, T. Proc. Asia Display, ’95, 1995; p 269.(380) Tasch, S.; Brandstatter, C.; Meghdadi, F.; Liesing, G.; Froyer,

G.; Athouel, L. Adv. Mater. 1997, 9, 33.(381) Nakamura, H.; Hosokawa, C.; Kusumoto, T. Inorganic and

Organic Electroluminescence/EL 96, Berlin, 1996; p 95.(382) Yoshida, M.; Fujii, A.; Ohmori, Y.; Yoshino, K. Jpn. J. Appl. Phys.

1996, 35, L397.(383) Bulovic, V.; Gu, G.; Burrows, P. E.; Thompson, M. E.; Forrest,

S. R. Nature 1996, 380, 29.(384) Burrows, P. E.; Gu, G.; Bulovic, V.; Forrest, S. R.; Thompson,

M. E. Device Research Conference, Santa Barbara, 1996.(385) Shen, Z.; Burrows, P. E.; Bulovic, V.; Forrest, S. R.; Thompson,

M. E. Science 1997, 276, 2009.(386) Takada, N.; Tsutsui, T.; Saito, S. Appl. Phys. Lett. 1993, 63, 2032.(387) Takada, N.; Tsutsui, T.; Saito, S. Jpn. J. Appl. Phys. 1994, 33,

L863.(388) Dodabalapur, A.; Rothberg, L. J.; Miller, T. M. Appl. Phys. Lett.

1994, 65, 2308.(389) Dodabalapur, A.; Rothberg, L. J.; Miller, T. M.; Kwock, E. W.

Appl. Phys. Lett. 1994, 64, 2486.(390) Lidzey, D. G.; Pate, M. A.; Wittaker, D. M.; Bradley, D. D. C.;

Weaver, M. S.; Fisher, T. A.; Skolnick, M. S. Chem. Phys. Lett.1996, 263, 655.

(391) Cimrova, V.; Neher, D. J. Appl. Phys. 1996, 79, 3299.(392) Jordan, R. H.; Rothberg, L. J.; Dodabalapur, A.; Slusher, R. E.

Appl. Phys. Lett. 1996, 69, 1997.(393) Dodabalapur, A.; Rothberg, L. J.; Jordan, R. H.; Miller, T. M.;

Slusher, R. E.; Phillips, J. M. J. Appl. Phys. 1996, 80, 6954.(394) Marks, R. N.; Biscarini, F.; Zamboni, R.; Taliani, C. Europhys.

Lett. 1995, 32, 523.

(395) Era, M.; Tsutsui, T.; Saito, S. Appl. Phys. Lett. 1995, 67, 2436.(396) Gu, G.; Shen, Z.; Burrows, P. E.; Forrest, S. R. Adv. Mater. 1997,

9, 725.(397) Gustaffson, G.; Treacy, G. M.; Cao, Y.; Klavertter, F.; Colaneri,

N.; Heeger, A. J. Synth. Met. 1993, 57, 4123.(398) Gymer, R. W. Endeavor 1996, 20, 115.(399) Tsutsui, T. Mater. Res. Soc. Bull. 1997, 22, 39.(400) Burrows, P. E.; Forrest, S. R.; Thompson, M. E. Curr. Opinion

Sol. State Mater. Sci. 1997, 2, 236.(401) Wakimoto, T.; Murayama, R.; Nagayama, K.; Okuda, Y.; Nakada,

H.; Tohma, T. Soc. Inf. Disp. Dig. 1996, 849.(402) Lovinger, A. J.; Davis, D. D.; Ruel, R.; Torsi, L.; Dodabalapur,

A.; Katz, H. E. J. Mater. Res. 1995, 10, 2958.(403) Torsi, L.; Dodabalapur, A.; Katz, H. E. J. Appl. Phys. 1995, 78,

1088.(404) Gundlach, D. J.; Lin, Y. Y.; Jackson, T. N. IEEE Electron. Dev.

Lett. 1997, 18, 87.(405) deLeeuw, D. M.; Brown, A. R.; Matters, M.; Chmil, K.; Hart, C.

M. Mat. Res. Soc. Spring Mtg., San Francisco, 1997; p PaperH3.5, 144.

(406) Dodabalapur, A.; Laquindanum, J.; Katz, H. E.; Bao, Z. Appl.Phys. Lett. 1996, 69, 4227.

(407) Guillaud, G.; Sadoun, M. A.; Maitrot, M.; Simon, J.; Bouvet, M.Chem. Phys. Lett. 1990, 167, 503.

(408) Hoshimono, K.; Fujimori, S.; Fujita, S.; Fujita, S. Jpn. J. Appl.Phys. Part 2 1993, 32, L1070.

(409) Lin, Y. Y.; Gundlach, D. J.; Nelson, S. F.; Jackson, T. N. 55thAnn. Dev. Res. Conf., Ft. Collins, CO, 1997; p 60.

(410) Blom, P. W. M.; Jong, M. J. M. d.; vanMunster, M. G. Phys. Rev.B 1997, 55, R656.

(411) Severt, B.; Horowitz, G.; Ries, S.; Lagorse, O.; Alnot, P.; Yassar,A.; Deloffre, F.; Srivastava, P.; Hajlaoui, R.; Lang, P.; Garnier,F. Chem. Mater. 1994, 6, 1809.

(412) Narioka, S.; Ishii, H.; Edamatsu, K.; Kamiya, K.; Hasegawa, S.;Ohta, T.; Ueno, N.; Seki, K. Phys. Rev. B 1995, 52, 217.

(413) Torsi, L.; Dodabalapur, A.; Lovinger, A. J.; Katz, H. E.; Ruel,R.; Davis, D. D.; Baldwin, K. W. Chem. Mater. 1995, 7, 2247.

(414) Dodabalapur, A.; Katz, H. E.; Torsi, L.; Haddon, R. C. Appl. Phys.Lett. 1996, 68, 1108.

(415) Ohno, T. R.; Chen, Y.; Harvey, S. E.; Kroll, G. H.; Weaver, J.H.; Haufler, R. E.; Smalley, R. E. Phys. Rev. B 1991, 44, 13747.

(416) Hosokawa, C. Appl. Phys. Lett. 1993, 62, 3238.(417) Sugiura, H.; Iga, R.; Yamada, T. J. Cryst. Growth 1992, 120,

389.(418) Forrest, S. R.; Burrows, P. E.; Haskal, E. I.; Zhang, Y. Mat. Res.

Soc. Ann. Mtg., Boston, MA, 1994; p 37.(419) Hayashi, K.; Kawato, S.; Fujii, Y.; Horiuchi, T.; Matsushige, K.

Appl. Phys. Lett. 1997, 70, 1384.(420) Tatsuura, S.; Sotoyama, W.; Motoyoshi, K.; Matsuura, A.;

Hayano, T.; Yoshimura, T. Appl. Phys. Lett. 1993, 62, 2182.(421) Tatsuura, S.; Sotoyama, W.; Yoshimura, T. Appl. Phys. Lett.

1992, 60, 1661.(422) Yoshida, Y.; Horiuchi, T.; Matsushige, K. Jpn. J. Appl. Phys.

1993, 32, 1248.(423) Kubono, A.; Kitoh, T.; Kajikawa, K.; Umemoto, S.; Takezoe, H.;

Fukuda, A.; Okui, N. Jpn. J. Appl. Phys. 1992, 31, L1195.(424) Takeno, A.; Okui, N.; Kitoh, T.; Muraoka, M.; Umemoto, S.;

Sakai, T. Thin Solid Films 1991, 202, 213.(425) Takeno, A.; Okui, N.; Kitoh, T.; Muraoka, M.; Umemoto, S.;

Sakai, T. Thin Solid Films 1991, 202, 205.(426) Perry, J. W.; Marder, S. R.; Perry, K. J.; Sleva, E. T. Proc. SPIE

1991, 1560, 302.(427) Kubono, A.; Higuchi, H.; Umemoto, S.; Okui., N. Thin Solid

Films 1993, 232, 256.(428) GaInAsP Alloy Semiconductors; Pearsall, T. P., Ed.; John

Wiley: New York, 1982.(429) Forrest, S. R.; Burrows, P. E.; Stroustrup, A.; Strickland, D.;

Ban, V. S. Appl. Phys. Lett. 1996, 68, 1326-1328.(430) Burrows, P. E.; Forrest, S. R.; Sapochak, L. S.; Schwartz, J.;

Buma, T.; Ban, V. S.; Forrest, J. L. J. Cryst. Growth 1995, 156,91.

(431) Tsang, W. T. J. Cryst. Growth 1990, 105, 1.(432) Tsang, W. T. J. Cryst. Growth 1992, 120, 1.(433) Feller, M. B.; Chen, W.; Shen, Y. R. Phys. Rev. A 1991, 43, 6778.(434) Zhuang, X.; Marrucci, L.; Shen, Y. R. Phys. Rev. Lett. 1994, 73,

1513.(435) Tillman, N.; Ulman, A.; Penner, T. L. Langmuir 1989, 5, 101.(436) Wu, S. Polymer Interface and Adhesion; Marcel Dekker: New

York, 1982.(437) Forrest, S. R. Optoelectronics in Japan and the United States;

Loyola College: Baltimore, 1996.(438) Kuwata-Gonokami, M.; Jordan, R. H.; Dodabalapur, A.; Katz,

H. E.; Schilling, M. L.; Slusher, R. E. Opt. Lett. 1995, 20, 2093.(439) Tessler, N.; Denton, D. J.; Friend, R. H. Nature 1996, 382, 695.(440) Hide, F.; Schwartz, B. J.; Diaz-Garcia, M. A.; Heeger, A. J. Chem.

Phys. Lett. 1996, 256, 424.(441) Hide, F.; Diaz-Garcia, M. A.; Schwartz, B. J.; Andersson, M. R.;

Pei, Q.; Heeger, A. J. Science 1996, 273, 1833.

Ultrathin Organic Films Grown by OMBD Chemical Reviews, 1997, Vol. 97, No. 6 1895

Page 104: Hongik - Ultrathin Organic Films Grown by Organic Molecular …huniv.hongik.ac.kr/~ekim/Ref_OS/126_127.pdf · 2004. 3. 5. · Ultrathin Organic Films Grown by Organic Molecular Beam

(442) Schwartz, B. J.; Hide, F.; Andersson, M. R.; Heeger, A. J. Chem.Phys. Lett. 1997, 265, 327.

(443) Kozlov, V. G.; Burrows, P. E.; Forrest, S. R. Conf. on Lasers andElectrooptics (CLEO ‘97), Baltimore, MD, 1997; p CPD-18.Kozlov, V. G.; Bulovic, V.; Burrows, P. E.; Forrest, S. R. Nature,in press.

(444) Soffer, B. H.; McFarland, B. B. Appl. Phys. Lett. 1967, 10, 266.(445) Kogelnik, H.; Shank, C. V. Appl. Phys. Lett. 1971, 18, 152.(446) Karl, N. Phys. Status Solidi 1972, 13, 651.

(447) Hermes, R. E.; Allik, T. H.; Chandra, S.; Hutchinson, J. A. Appl.Phys. Lett. 1993, 63, 877.

(448) Diaz-Garcia, M. A.; Hide, F.; Schwartz, B. J.; McGehee, M. D.;Andersson, M. R.; Heeger, A. J. Appl. Phys. Lett. 1997, 70, 3191.

(449) Schafer, F. P. Dye Lasers; Springer-Verlag: Berlin, 1977.(450) Haskal, E. I. University of Southern California, 1994.

CR941014O

1896 Chemical Reviews, 1997, Vol. 97, No. 6 Forrest