Top Banner
STUDY OF RESONANCE LIGHT SCATTERING FOR REMOTE OPTICAL PROBING by C.M. Penney (Principal Investigator), W. W. Morey, R. L. St. Peters, S. D. Silverstein, M. Lapp, and D. R. White Prepared under Contract No. TSTAS 1-11624 by General Electric Corporate Research and Development P.O. Box 8 Schenectady, New York 12301 for NATIONAL AERONAUTICS AND SPACE ADMINISTRATION LANGLEY RESEARCH CENTER HAMPTON, VIRGINIA 23365 September 1973 SRD-73-125 https://ntrs.nasa.gov/search.jsp?R=19740012203 2020-06-05T05:54:48+00:00Z
139

study of resonance light scattering for remote optical probing

Sep 11, 2021

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: study of resonance light scattering for remote optical probing

STUDY OF RESONANCE LIGHT SCATTERINGFOR REMOTE OPTICAL PROBING

byC.M. Penney (Principal Investigator), W. W. Morey, R. L. St. Peters,

S. D. Silverstein, M. Lapp, and D. R. White

Prepared under Contract No. TSTAS 1-11624

by

General Electric Corporate Research and DevelopmentP.O. Box 8

Schenectady, New York 12301

for

NATIONAL AERONAUTICS AND SPACE ADMINISTRATIONLANGLEY RESEARCH CENTER

HAMPTON, VIRGINIA 23365

September 1973

SRD-73-125

https://ntrs.nasa.gov/search.jsp?R=19740012203 2020-06-05T05:54:48+00:00Z

Page 2: study of resonance light scattering for remote optical probing

TABLE OF CONTENTS

Page

LIST OF ILLUSTRATIONS v

SUMMARY 1

I. INTRODUCTION 2

Scattering from Molecules 5Luminescence from Molecules : 5Scattering and Luminescence from Particles 7Optical Diversion Processes and Gas Probing 7

II. THEORY 11

Near-Resonance Raman Scattering 11Simple Quantum Mechanical Calculations 23Near Resonance Raman Scattering From Atoms 28Near Resonance Raman Scattering From Molecules 28Fluorescence 40

III. EXPERIMENT - 47

Double Monochromator Facility 47Experiments with NO2 51Experiments with Iodine Vapor 55Measurement of the Ozone Cross Section at Visible Light

Wavelengths 60Spectrometer-Dye Laser Facility 61Spectrometer and Related Optics 68Detection and Time Analyzing Electronics 69Signal Counting Statistics 71Absolute N2 Vibrational Raman Scattering Cross

Section for Incident Light at SOOOnm 72The O2 Vibrational Raman Cross Section for Incident Light

at 300 nm 75Laser-Excited Fluorescence from SO2 75Low Pressure SO2 Fluorescence 76

IV. MODEL CALCULATIONS 92

Ground-Based Lower Atmosphere Probe 92Ground-Based Upper Atmosphere Probe 93Airborne Measurements 95Measurement of Upper Atmosphere Constituents from

a Satellite '• 98

V. CONCLUSIONS -- 102

Near Resonance Scattering 102Scattering Following Excitation in Dissociative Continuum 103Fluorescence 104Experimental Results 105Model Calculations 105

iii

Page 3: study of resonance light scattering for remote optical probing

TABLE OF CONTENTS (Continued)

Page

APPENDIX A 107

APPENDIX B 109

APPENDIX C 111

APPENDIX D 121

REFERENCES 125

IV

Page 4: study of resonance light scattering for remote optical probing

LIST OF ILLUSTRATIONS

Figure Page

1 LIDAR single ended optical probe 3

2 . Bi-axial optical probe 4

3 Level diagram illustrating various optical diversionprocesses 6

4 Functional plot for the "near-but-off-resonance"case 15

5 Functional plot for the partial overlap case 15

6 Plots of the time-dependence of detector responses-- 21

7 Level diagrams for selected metal atoms 29

8 Re-emission cross sections for aluminum 30

9 Ultraviolet absorbtion bands of O3, SO2, and NO 31

10 Scattering geometry for definition of depolarization - 43

11 Block diagram of the Double MonochromatorFacility 48

12 Spectral distribution of NO and NO2 re-emission 53

13 Typical spectral trace for NO2 53

14 Spectral distribution of NO2 Vj-line fluorescence 55

15 Absorption of I2 vapor as a function of laser wave-length near 514. 5 nm 57

16 Antistokes fluorescence from iodine • 57

17 Normalized collision and main peak intensities as afunction of separation from resonance 59

18 Tunable dye laser and optics for scatteringexperiments 62

19 Flowing dye cell for N2 Laser Pumped-Dye Laser - - 64

20 Block diagram of the Monochromator-Dye LaserFacility . 70

21 Time dependence of Rayleigh/Mie scattering 72

22 Relationship between true and indicated count ratesfor the MDLF - 73

23 Typical signal obtained for vibrational Ramanscattering 74

Page 5: study of resonance light scattering for remote optical probing

Figure Page

24 Fluorescence spectrum of low pressure SO2

excited at 299. 98 nm 77

25 Fluorescence spectrum of low pressure SO2

excited at 299. 90 nm 79

26 Tuned Laser Fluorescence spectrum of SO2 • 80

27 Comparison of TLF and absorption spectra 82

28 Time dependence of v:-shifted re-emission from SO2 - 83

29 Stern-Volmer plot of SO2 self-quenching 85

30 Fluorescence spectrum of SO2 in 700 Torr N2 - •• 87

31 Intensity of Vj-line fluorescence from SO2 asafunctionof background air pressure 88

32 Stern-Volmer plot of SO2 quenching by air 90

33 Tuned laser fluorescence spectrum of SO2 Vj-line inair 91

34 Minimum measurable concentrations of atoms andOH radical for satellite probe 99

VI

Page 6: study of resonance light scattering for remote optical probing

STUDY OF RESONANCE LIGHT SCATTERINGFOR REMOTE OPTICAL PROBING

C.M. Penney (Principal Investigator), W.W. Morey, R. L. St. Peters,S. D. Silverstein, M. Lapp, andD.R. White

SUMMARY

The.objective of the work described in this report is to investigate en-hanced scattering and fluorescence processes in the visible and uv which willenable improved remote measurements of gas properties.

The theoretical relationship between scattering and fluorescence froman isolated molecule in the approach to resonance is examined through analysisof the time dependence of re-emitted light following excitation by pulsed incidentlight. Quantitative estimates are developed for the relative and absoluteintensities of fluorescence and resonance scattering. New results are ob-tained for depolarization of scattering excited by light at wavelengths withina dissociative continuum.

The experimental work was performed in two separate facilities. One ofthese utilizes argon and krypton lasers, single moded by a tilted etalon, anda 3/4 meter double monochromator. This facility was used to determineproperties of the re-emission from NO2, I2 and O3 excited by visible light.

The second facility involves a narrow-line dye laser, and a 3/4 metersingle monochromator. The dye laser produces pulsed light with 5 nsec pulseduration and 0. 005 nm spectral width. The spectrometer and associatedelectronics provide both time (6 nsec) and spectral (0. 02 nm) resolution ofthe re-emitted light. This facility was used to measure the absolute vibra-tional Raman scattering cross sections of N2 and O2 at 300 nm, and to ex-amine in detail the fluorescence from SO2 excited at various wavelengthsnear 300 nm. It was found that both N2 and O2 cross sections are severaltimes larger than predicted by a dAscatteP4 extrapolation from the visible.Sharp line fluorescence from SO2 in air near STP was observed to be 104

time stronger than the N2 vibrational Raman scattering. Comparable orlarger cross sections for O3 and numerous atomic species are estimated.

Use of the strong scattering and fluorescence processes in variousmeasurement situations is considered. Among the remote, range-resolvedmeasurements that appear practical are: ambient SO2 and O3 in the loweratmosphere from the ground; O3 in the stratosphere from the ground; andnumerous atomic species at altitudes above 50 km from a satellite.

Page 7: study of resonance light scattering for remote optical probing

I. INTRODUCTION

The objective of the work described in this report is to identify and in-vestigate optical processes in the visible and ultraviolet that can expand thecapabilities of remote, spatially resolved measurements of gas properties.The optical processes considered here are those that divert light from anincident beam, such that some of the diverted light returns to a collector.Intensity, spectral and/or polarization analysis of the collected light providesinformation about the gas volume from which it came.

One type of probe based on this principle is the single-ended LIDARsystem, illustrated in Fig. 1. This type of probe has already been employedwith considerable success for ground- and air-based studies of the atmosphere(refs. 1 and 2) and shows good potential for satellite-based studies. A secondexample is the bi-axial system shown in Fig. 2, which can be used to obtainprecise three-dimensional resolution in relatively small regions such as awind tunnel (ref. 3).

There are a number of optical diversion processes upon which suchprobing systems can be based. Ideally, the observed process should havethe following characteristics:

(a) It should be effectively instantaneous. This characteristic is parti-cularly desirable for LIDAR applications, where ranging is accom-plished through timing. In practice, it is often preferable that theobserved process introduce less than 10~8 second time uncertainty.

(b) The intensity of the diverted light per molecule should be independentof gas pressure and constituency.

(c) The returned light should be characteristic of the observed mole-cules and their states of excitation.

(d) The process should be strong enough for easy observation, but not sostrong as to prevent light transmission to the measurement point.

We shall consider several optical processes which meet these criteriato varying degree. The processes can be divided into two types: scatteringand fluorescence. Within the title of this report, the work " scattering" isused to denote both of these processes generically. However, subsequently,we will use this term exclusively to designate those processes which areeffectively instantaneous and insensitive to collision effects such as relaxation,collisional depolarization and quenching. On the other hand, we shall use theterm fluorescence to designate processes which are intrinsically " slow" , dis-playing a time uncertainty at low gas pressures which is typically greaterthan 10~8 sec, and a corresponding sensitivity to collision effects at highpressure.

Page 8: study of resonance light scattering for remote optical probing

CO -M

O. O

Page 9: study of resonance light scattering for remote optical probing

INCIDENT LIGHT BEAM

\MEASUREMENT

REGION

COLLECTIONLENS

FILTER SYSTEMOR SPECTROMETER

Figure 2 Bi-axial optical probe. This configurationprovides convenient three-dimensional resolution atclose range.

Although fluorescence is often thought of as a process following actualexcitation into an excited state, the distinction between fluorescence andscattering does not always correlate with separation from resonance. Thereare situations to be considered in which excitation in a strong absorptionregion produces scattering. This type of process is often called resonancescattering (ref. 4). However, the use of such terms in the literature extendsto processes which could also be regarded legitimately as fluorescence, pro-ducing a somewhat confusing situation. In Appendix A we describe severaldifferent criteria for distinguishing between scattering and fluorescence thathave appeared in the literature. The present choice is based on two consider-ations: First, scattering has been regarded by many as an "instantaneous"process, in which case there is not time for collisions to occur. Second, thischoice is convenient for gas probe applications, because under conditions ofpresent interest scattering satisfies the first two "ideal" properties we havelisted for a probe, whereas fluorescence often fails one or both of thesecriteria.

Because nature often fails to accommodate idealism, ultimately we willconsider both scattering and fluorescence processes in some detail. Thus atthis point we present a descriptive list of some of the general types of optical

Page 10: study of resonance light scattering for remote optical probing

diversion processes that are potentially useful for gas probing. The listserves to introduce terminology, and to provide an initial orientation con-cerning these processes. The general types of transitions involved in themolecular processes included in this list are illustrated by level diagrams inFig. 3.

Scattering from Molecules i

Rayleigh scattering. - In this process light is scattered from a moleculewhithout exchanging any energy with the internal states of the molecule. Onlythe very small amount of energy necessary to conserve momentum betweenthe light photon and scattering molecules is exchanged (Doppler-Brillouineffect). Therefore the scattered light has nearly the same spectral distribu-tion as the incident light, and it is very difficult to determine the relativeconcentration of different molecules from this spectral distribution. Temp-erature information is accessible from the spectral distribution, but, becauseof the small amounts of energy exchanged, it requires rather delicatemeasurements.

Raman scattering. - This process involves an exchange of a significantamount of energy between the scattered light photon and the scattering mole-cule. Consequently, the scattered light undergoes substantial shifts inwavelength. The resulting bands of scattered light are characteristic of theparticular molecule. The intensity of a band is proportional to the number ofmolecules in the particular initial state leading to that band. Thus Ramanscattering provides direct information about both the constituency and excitedstate populations of molecules in a system.

Elastic light scattering by particles. - In this process, which is oftencalled Mie scattering (appropriate for spherical homogeneous particles), lightis scattered from particles suspended in the gas. The light undergoes onlythe (usually) minute spectral shift arising from the particle motion. However,the dependence of the scattering on incident wavelength, on polarizations of in-cident and scattered light, and on scattering angles can be used to obtain someinformation about the particle distribution.

Luminescence from Molecules

Fluorescence. - This process can be initiated by absorption of light at awavelength within a particular absorption line or band of a molecule. Subse-quently, the molecule may re-emit, light in a transition from the excited stateto the original or a different lower state. The probability of re-emission perunit time y is usually constant such that at very low pressures the fluorescence,intensity displays an exponential decay curve with time constant 1/y. However,before a molecule can de-excite by emitting light, the excitation may be

Page 11: study of resonance light scattering for remote optical probing

1

1 1

1 \i

1111

T i

RAYLEI6H RAMAN FLUORESCENCE FLUORESCENCESCATTERING SCATTERING AND

PHOSPHORESCENCE

Figure 3 Level diagram illustrating various optical diversion processesin molecules.

Page 12: study of resonance light scattering for remote optical probing

altered by some non-radiative process . For example, the molecule mayundergo collisions with other molecules. These collisions can have severaleffects. One effect is that they can cause the molecule to lose "memory" ofthe direction and polarization of the incident beam, increasing the depolariza-tion of the fluorescence and decreasing the anisotropy of its angular distribu-tion. Collisions in which small amounts of energy are exchanged redistributethe excitation among various upper molecular states, broadening the spectral

v distribution of the fluorescence. In other collisions all or most of the excita-tion energy is degraded in a non-radiative process. For example, it may beconverted to kinetic energy. The latter phenomenon reduces the fluorescenceintensity and is, therefore, referred to as quenching. Quenching also shortensthe decay time of re-emission, as mentioned previously. This effect isadvantageous for time-of-flight ranging. However, in general, collisionprocesses render the fluorescence spectral distribution, polarization and in-tensity per molecule dependent upon the gas constituency and total pressure.Therefore, the effect of these processes must be known or calculated beforequantitative data can be obtained from fluorescence. In particular, if thecollision effects are dominated by a species of variable concentration, theconcentration of the quenching species must be measured separately.

Although collision effects tend to smooth out the spectral detail in fluores-cence, nevertheless enough detail is often retained even at atmosphericpressure to allow identification of the fluorescing species.

Phosphorescence. - This process involves excitation to an upper state,followed by a transition, often collision-induced, to a state from which radia-tive decay is forbidden. Phosphorescence is composed of the slow forbiddendecay, and of fast decay following a second transition, again often collision-induced, back to the states with allowed transitions. Since phosphorescenceinvolves a more detailed collision history than fluorescence, and is often sloweven at high pressure, it is rarely used in gas probing.

Scattering and Luminescence from Particles

In addition to the processes mentioned in the list above, particles cancontribute Raman scattering, fluorescence and phosphorescence. Amongthese, Raman scattering is very weak, but fluorescence seems likely toprovide useful measurements of particle properties. However, this subjectwill not be addressed specifically in this paper.

Optical Diversion Processes and Gas Probing

Raman .scattering. - Among the processes discussed above, only Ramanscattering (RS) completely satisfies the first three "ideal" properties for anoptical probe; i. e., it is instantaneous, not sensitive to quenching, and charac-teristic of the type and excitation of the scattering molecules. Furthermore,RS does not require a tunable laser, and data analysis to obtain desired

Page 13: study of resonance light scattering for remote optical probing

information about the gas is relatively simple. However there is one substan-tial disadvantage of ordinary RS - it is relatively weak. For example, atypical RS cross section for scattering of visible light is 10~30 cm2/sr permolecule (ref. 5 .) . In contrast, a typical Rayleigh cross section is about 1000times larger (ref. 6), and a fluorescence cross section, about 1010 to 1016

times larger (in the absence of quenching). Because RS is so weak, RS probeswith high sensitivity at long range require large lasers and collector lenses.In fact, limitations of size, weight, expense, safety and/or present technologyare often reached all too soon in such applications. That point will be examinedin some detail in Chapter IV. However at the present time it serves to intro-duce the question of major interest in this .report; namely, can we find pro-cesses which are much stronger than ordinary RS, yet share to a useful extentits desirable characteristics for optical probing?

Enhanced scattering upon approach to resonance. - Indeed, the possibilityof strongly enhanced RS itself is suggested by the resonance-like structureof the basic quantum expression for scattering. We will argue that thisequation, which is presented in Chapter II, describes both scattering andfluorescence, at least in simple systems represented by a two or threelevel quantum model. In this case the resonance peaks produce fluorescence,and not scattering. As the incident light frequency approaches a strong, . ' ' .isolated absorption line, the character of the diverted light will undergo agradual transition from that of scattering (effectively instantaneous andnot subject to quenching) into fluorescence. Furthermore, the wavelengthdependence of this transition may depend sensitively on detailed line broadeningmechanisms which are operative in the experimental situation. (In fact, ob-servation of the scattering-to-fluorescence transition should provide newinformation about line broadening in the far wings of the absorption line.)However, qualitative theoretical arguments suggest that the intensity of RScan be substantially enhanced in the approach to a strong isolated absorptionline or band before the transition to fluorescence-like properties occurs. Thistype of scattering, which has been called pre-resonance or near-resonanceRaman scattering, has attracted interest among those who use RS as a gasprobe; therefore, it has received substantial attention in the work reportedhere.

Scattering following excitation into a dissociative continuum. - In thisprocess a molecule is excited by light with sufficient photon energy to causeit to dissociate. However, there is a distinct probability that the moleculewill de-excite by re-emitting light before it dissociates. This light displaysscattering-like properties because typically dissociation will occur so quicklythat the excited molecule has little chance to experience a collision. Con-sequently the re-emitted light appears "instantaneously" and is insensitive to

8

Page 14: study of resonance light scattering for remote optical probing

collisions and associated quenching, etc. This process has been identifiedand studied experimentally by Holzer, Murphy and Bernstein (ref. 4) in halogenmolecules. Aspects of it have been examined theoretically by Berjot et al.(ref. 7, 8). In particular, moderately strong enhancement of scattering fromI2 has been found under excitation at wavelengths shorter than the dissociationlimit near 510 nm (ref. 4).

Scattering over quenched fluorescence. - This process has been discussedby Berjot, Jacon and Bernard (ref. 9, 10) for situations when fluorescenceis strongly quenched, leaving a remnant of scattering from intermediate statesnot in resonance. In this case the intensity of diverted light per absorbingmolecule should decrease in propagation to background gas pressure until a" floor" of scattering is reached. Simultaneously the depolarization shouldgo through a characteristic variation. Berjot et al. have examined thisphenomenon theoretically and shown favorable comparison to experimentswith lz in high pressure (up to 40 atml) argon. However, in Chapter III weshall show that their results may arise at least partially from an alternativemechanism.

Quenching balanced by increased absorption due to line broadening. - Therate of fluorescence per molecule Rp can be expressed as the product of arate RA of absorption and a probability PE of re-emission before quenching(ref. 11). Although Pg will decrease as pressure increases, R^ can increaseat a sufficient rate to balance this decrease (or even over-balance it), if theexciting wavelength is on a wing of the absorption line. The increase of R^can be explained as follows: Both theory and experiment have demonstratedthat the area under an absorption line (integral of absorption cross sectionover wavelength or frequency remains very nearly constant as collisionsbroaden the line (ref. 12). Therefore, the absorption must decrease online center, but increase in the wings. This effect has been studied in I2

vapor by Fouche et al. (ref. 13), and St. Peters et al. (ref. 14). The formergroup first demonstrated that an enhancement on the order of 106 can be ob-tained at a separation of about 0.1 cm"1 from line center, and that this strongre-emission is fairly insensitive to pressure, varying by a factor of < 10over the range from a few torr to 700 torr. However, there is a strongcollision-induced continuum in the re-emission spectral distribution at thehigher pressures studied. Therefore, we consider this process to be a formof fluorescence, albeit a potentially useful form.

General fluorescence. - Fluorescence is much stronger than ordinary RS,typically by factors of lO^-lO^at low pressure, as mentioned previously.However, it is often more difficult to decipher density and temperature in-formation from measurements of fluorescence intensity, because of resonanceabsorption of the incident beani (which is sensitive,to the fine detail of linebroadening), because of quenching and, at low pressure, because of the delayin re-emission which complicates time-of-flight ranging. Furthermore, ifnarrow line absorption is involved, then efficient use of the incident beam

Page 15: study of resonance light scattering for remote optical probing

requires that its spectrum be commensurately narrow and precisely controlled.

Nevertheless, we believe that there are many situations in which fluores-cence can be used effectively to probe the atmosphere. One example whichhas already been implemented is the particularly successful observationof Na and K vapor in the upper atmosphere (refs. 15, 16). These measure-ments, which have been made from distances on the order of 80 km, aresuccessful largely because the observed atoms are concentrated in a relativelythin layer and their absorption cross sections are very large, producing decaytimes which are too fast for substantial quenching at that altitude.

However, there is at least one other favorable situation for fluorescenceprobing of the atmosphere, which has relatively wide applicability. Thissituation satisfies the following conditions:

1. The species whose measurement is sought exists at low concentration*in a carrier gas which dominates line broadening and quenching.

2. The pressure of the carrier gas is known with sufficient accuracy tomake line broadening and quenching corrections, and it is high enoughto shorten the re-emission delay of observed fluorescence to the pointwhere time-of-flight ranging can be used.

3. The quenched fluorescence of the target species is much strongerthan ordinary RS and has a characteristic spectral distribution whichallows its identification.

Under these conditions, the fluorescence will be nearly proportional to theconcentration of the observed species, and a single-ended LIDAR instrumentcan be used to observe it with time-of-flight ranging. Moreover, if theabsorption bands are fairly wide and smooth, then the requirements on thelinewidth and spectral stability of the laser source are not critical. Theseconditions are satisfied for NOg and SOg. Relevant characteristics offluorescence from these gases will be described subsequently in thisreport.

The two fields of enhanced scattering near or in absorption regions, andlaser excited fluorescence have experienced rapid acceleration in the last fewyears. For example, in this period over 100 papers have been published onresonance Raman scattering and closely related subjects. In this introductionwe have endeavored to give an overview of those developments which pertaindirectly to LIDAR probing of the atmosphere. In subsequent chapters, we willexplain our own work in more detail, and place it into context with the workof other groups and with ultimate desired applications.

10

Page 16: study of resonance light scattering for remote optical probing

IL THEORY

In this chapter we present theoretical analyses of various aspects of thescattering and fluorescence processes described in the previous chapter.Work performed under this contract will be described in detail and placed intocontext by brief descriptions of related work performed here under othersupport, and elsewhere.

Near-Resonance Raman Scattering

Classical Analysis of Scattering. - Here we develop a relatively simpleclassical model which represents Rayleigh and Raman scattering, as well asfluorescence. Although this classical analysis is not quantitatively correct,it provides a significant qualitative picture of a transition from scattering tofluorescence in the approach to resonance. This picture is confirmed inlimiting cases by a detailed quantum-mechanical calculation which will bedescribed subsequently.

Scattering has been characterized by an effectively instantaneous timeresponse, and insensitivity to collision effects. In the case of pulse excita-tion, these two properties would seem to be closely related, for if theresponse is instantaneous there is not time for a molecule to experience acollision. Initially, we will use the classical model to investigate the timedependence of re-emission following pulsed excitation. Subsequently, weshall infer the significance of collision effects insofar as possible.

We consider a charged oscillator whose response represents that of amolecule. The alternating electric field of an incident light beam will setthis oscillator in motion, such that it radiates a second alternating field whichrepresents the diverted radiation. This field alternates at the same frequencyas the incident radiation (for a stationary oscillator) and thus constitutesRayleigh scattering or resonance fluorescence. However, if the oscillatoris, say, rotating, this motion will modulate the properties of the field.Thereupon side bands will be produced in the radiation at a separation fromthe incident light frequency which is equal to the rotation frequency. Theseside bands represent rotational Raman scattering or fluorescence. Othertypes of Raman scattering and fluorescence (vibrational, vibrational-rotational, etc. ) can likewise be represented through this model.

We wish to discover what happens as the frequency of incident light isbrought near resonance with the oscillator; in particular we wish to ascertainthe amplitude and time dependence of the diverted radiation when the incidentlight is in the form of a short pulse. To do so, we represent the incidentlight at the oscillator in terms of its electric field

11

Page 17: study of resonance light scattering for remote optical probing

It = ~ E? (t) (1)

and introduce the Fourier transform E^uu) of the field such that

r +•Ei(t) = / ^ E! (uu)e 1U)t (2)

The response of the charged oscillator is described by the differentialequation

X(t) + rf»X(t) + ^^ X (t) = - — Ei (t) (3)3mc3 m

where X(t) is the separation between the charges, e is the charge magnitude,m is the reduced mass associated with the charges, and 0 is the (angular)resonance frequency.

«Introducing the Fourier transform of X(t) through

X(t) = f f X(u)e+ia)t (4)

— CO

and taking the Fourier transform of Eq. (3), we obtain

(5)m uu -Cr-

Where

Y = - - 3-3mc

The far field of the scattered radiation at a distance R from the oscillator andin a direction perpendicular to its oscillation is given by

E2(R,t) = --2-T X(t --) (7)Re2 V c/

or using Eq. (2)

Re

12

Page 18: study of resonance light scattering for remote optical probing

Here we consider only the scattered field at the same frequency as the incidentfield; i. e. , Rayleigh scattering. The frequency-shifted fields, arising sayfrom rotation of the oscillator (Raman scattering), can be expected to show thesame general time dependence.

Substituting for X(uj) from Eq. (5) we obtain

R duu oi2Ei(uj)e2TT S-

+iujt(9)

and therefore, the total scattered irradiance at a distance R is given by

i2

12 R't+ =

duj a)2E! (uj)e2n ..^-o2 (10)

The angle dependence of this irradiance, which we shall omit in this deriva-tion, is given by cos2ijf, where ty is the angle between the polarizations ofincident and observed re-emitted light.

Suppose that the scattered radiation is analyzed by a spectrometer (orfilter) with passband S(|uu|). The irradiance of scattered light within thispassband is

I2(R, t) =4TTR2\mc2 2TT

(ID

where we have omitted the time-of-flight delay R/c.

Equations (1), (2) and (10) relate the time dependence of the incident andobserved scattered radiatioa To examine this dependence for a simple casewe assume that the incident light is in the form of a coherent pulse withGaussian time distribution and pulse length I/A; i. e.

Ei(t) -t2A2/;cos u)i t (12)

The Fourier transform of this pulse is

(13)

13

Page 19: study of resonance light scattering for remote optical probing

where K is a normalization constant. It is convenient to express this constantin terms of the energy per unit area Jr in the incident pulse. Thus

a _ 4n3/2J1

cA

Substituting from Eq. (13) into Eq. (10), we obtain

I2(t) =

(14)

4rrR2 S(u>) uu2e"eiu)t + C.C.

u u - l r -(15)

where C. C., signifying the complex conjugate of the integral, arises from thesecond term in Eq. (13). It is apparent from this result that the time depen-dence of the scattered light is determined by an integral of the form

,duo 2 -(t«-uJi)2/2 A2

ore * ' . ,wot (16)

We have not been able to obtain a general closed form expression for this in-tegral. However its qualitative behavior can be inferred from a plot of partof the integrand. Furthermore, Eq. (15) can be solved for several limitingcases. Finally, numerical solutions can easily be obtained using a computer.We will discuss each of these approaches below.

If a wide, constant passband is assumed, S(uu) may be set equal to one, andthe integral in Eq. (15) becomes a Fourier transform of the function

F(ou) =uo2-f^-iuuY

(17)

Note that this function is a product of two strongly peaked functions, i. e. ,

U) and e- (ID-0)1 )2/2 A2

U) -\T-

A plot of |F(u))| is shown in Fig. 4 for the "near-but-off" resonance case.Except for a change in amplitude and slight shift in frequency, this functiondiffers but little from the Gaussian frequency distribution of the incident light.Therefore, in this case we expect the re-emission to follow the time depen-dence of the incident pulse; i. e. , it will be effectively instantaneous. This

14

Page 20: study of resonance light scattering for remote optical probing

(Ill

FREQUENCY

Figure-4 Plot of absolute values of the two peaked functions in F(iw), andtheir product, for the "near-but-off-resonance" case. The dashed linemostly to left represents abs^/fu^fft^-iuur)], that mostly to right representsexp[-(uu-u)1)2/2A2], and the solid line represents the normalized product.

FREQUENCY

Figure 5 Plot of absolute values of the two peaked functions in F(u)), andtheir product, for the "intermediate11 (partial overlap) case.

15

Page 21: study of resonance light scattering for remote optical probing

conclusion follows from the fact that in each case (incident and observedscattered light) the time dependence is determined by similar integrals of thefrequency dependent function.

In Fig. 5 we show an intermediate case, in which the spectral distributionof the incident radiation substantially overlaps the absorption line. In thiscase one component of the product closely resembles the Gaussian distribution,but a small second peak appears near the center of the absorption line. Herewe expect that the light in the spectral region of the Gaussian will appear asa short pulse, whereas the light centered in the spectral region of the smallerpeak will exhibit a slower, quasi-exponential decay. Thus, in this case bothscattering and fluorescence appear in the diverted light.

Finally, if the spectral distribution of the incident light is a broad peakcentered on a relatively narrow absorption line, corresponding to traditionalfluorescence, then the product has nearly the same shape as the absorptionline; therefore, in this case the re-emitted light will follow the expected,relatively slow exponential decay of fluorescence.

We may obtain approximate expressions for the magnitudes of fluorescenceand scattering for the case of near-but-off resonance by utilizing an approxi-mate form for F(uo). In this approximation, F(tu) is replaced by the sum oftwo functions, in each of which the wing of one peak is evaluated at the centerof the other; i. e. ,

, 2 -(u)-uuo)2/2 A2 rf -(n-uj0)2/2A2

SB« - - + -Ef— - : - - (18)ujo -fl -iujy uj2 -fi2-it«Y

This approximation can be quite good when the separation from resonance islarge compared to A, although it breaks down very near resonance. Inparticular, when the separation from resonance is large compared to both Aand y, and A > > y» the first term in F(uu) clearly leads to scattering and thesecond to fluorescence. We have just argued that both processes may existsimultaneously. It is possible to estimate the absolute magnitudes of each,and their ratio by substituting from Eq. (18) into Eq. (15). We obtain anexpression for Ijdw) which contains a square of the sum of two integrals, onearising from each term in F(iu). Cross products of these integrals producedin the squaring operation lead to contributions which "beat", oscillating at afrequency equal to the separation from resonance. This type of beating willbe washed out by collision effects and Doppler broadening in most cases,Thus, for present purposes, we. drop these terms, leaving two componentsto la; i. e.

16

Page 22: study of resonance light scattering for remote optical probing

where /:and

(19)

iurt+ C.C.

(20)

Continuing our assumption of a broad passband (S(u)) = 1), we can easilyevaluate these integrals, obtaining

and

/f?R2A

= 0, t < 0.

(21)

(22)

It is clear that I2S follows the time dependence of the incident pulse, whileI2f shows the exponential decay of fluorescence. Integrating over time weobtain the total energy (per unit area) in each component:

2

'as (23)

t2

/frR Ay \ me(24)

The results for the fluorescence components I2f(t) and J2f are of mostlyacademic interest, since typically these will be significant only near thelimit of applicability of the approximation, and there they are likely to bevery sensitive to inhomogeneous pressure and Doppler broadening, whoseeffects we have not included.

However, the results for the scattering component, I2g(t) and J2S are ofpractical interest, because they remain significant far from resonance wherethe approximation is very good. If Eq. (23) is multiplied by the angle-dependent factor appropriate to this simple model, cos2i|i, summed over

17

Page 23: study of resonance light scattering for remote optical probing

polarization and integrated over the area of the sphere at distance R, thetotal scattered energy is obtained in the form

(25)

Defining a scattering cross section

(26)

and employing the near-resonance approximation

= a

we obtain

(27)

(28)

It is interesting to compare Eq. (28) with the quantum mechanical resultpresented subsequently in Eq. (45). Except for numerical factors which arenear unity for strong transitions, they are the same.

So far we have discussed the time dependence of ^(t) through qualitativearguments and limiting cases. Now, to be more precise, we consider adirect numerical solution for kft) under conditions which are in the transitionzone between limiting cases. Starting with Eq. (15), we first put it into amore convenient form for computer integration.

Let

Then

J(UD) =

I2(t) =

cK2

nR2

V2

(29)

duj-

iuut (30)

neglecting unmeasurable high frequency components. It is convenient to takeexplicit account of a detector response in order to limit the necessary com-puter operations; therefore, let us define R(t, t1) to be the response of adetector at time t to a signal at time t1. Then the detector signal will be

18

Page 24: study of resonance light scattering for remote optical probing

Z(t) = I R(t, t ' JMt 'Jdt ' (31)

Substituting for Mt) from Eq. (30) and writing out the absolute square, weobtain

Z ( t ) = dt 'R(t , f ) - j * ( t t , ) J ( « ) ' ) e - (32)

Introducing a= u)1 -tu and inverting the order of integration, we may put thisequation into the form

Z(t) = J*(uu) J(uu+a) dt'R(t,t')e iat ' (33)4ir

Now suppose that the detector response is exponential; then

1 -ft-tR(t,t ' ) = -e v , t > t ' ; = 0, otherwise (34)

T

and, performing the integration over t', we obtain

00

1 1CXT

Z(t) = — T da I d(a J*^> J(uu+a) (35)

It is convenient to .define the correlation function

fG(a) = I diu J*(uu) J(uo+a) (36)

— 00,

Then note G(-a) = G#(oc). Using this property and the fact that the J's arenegligible for negative values of uu, we can write Eq. (35) in the form

S r00I 10.T

da -£- G(a)V (37)1+1CXT

19

Page 25: study of resonance light scattering for remote optical probing

where the symbol R denotes the real part of the integral.

In the computer calculation the correlation function G(a) is first cal-culated. This operation is facilitated by the fact that the J's are significantover only a small spectral region. Furthermore, large values of ado nothave to be considered because of the denominator in the a-integral. Aftercalculation of the correlation function, a fast Fourier transform routine isused to obtain Z(t) and this function is displayed by a digital plotter. Theresults obtained over a representative range of conditions are shown inFig. 6. The transition from delayed to instantaneous re-emission in thesimple case considered is quite evident.

A quantum mechanical analysis of the time dependence of re-emissionfrom a molecule following pulsed excitation has been developed by Seth D.Silverstein (see Appendix C). In this analysis the light pulse is representedby a propagating coherent wave packet of the type introduced by Glauber(ref. 17). The results confirm those of the classical approach describedabove.

Relevant conclusions can be drawn from these results along what mightbe considered as three levels of conjecture. First we consider pulsed ex-citation of a gas at sufficiently low pressure such that interactions betweenmolecules can be neglected. In this situation the distinction between scatter-ing and fluorescence concerns only the time dependence of the re-emission.The behavior of the classical model implies that "instantaneous" re-emission(i. e. , scattering) will be observed when the separation from resonance isseveral times larger than the widths of the excitation and absorption lines.We have not included the line shifts resulting from particle motion in thisanalysis. However, the resulting Doppler broadening falls off extremelyrapidly in the line wings, such that we may conclude that scattering will beobserved under the condition described above providing the absorption linewidth is considered to include the Doppler broadening.

At the second level of conjecture, we consider the effect of particleinteractions on the re-emission as the pressure is increased. At STP, themean time between collisions TC for small molecules is on the order of 10~10

sec,, whereas the duration of a collision is on the order of 10~n to 10~12 sec.If we may regard a molecule as "isolated" in between collisions, then theresults of the classical model apply. However, in this case the interactionbetween molecule and light must occur in a time short compared to TC, whichnecessitates that the spectral width of the exciting pulse be large in compari-son to the collision broadening width

"Y =* I/T'c ' c . •

20

Page 26: study of resonance light scattering for remote optical probing

G^ <u£ ra

•5 £ S

CQCD

Scfl

§CQ

ra(H

rt ri

o

CQ Gc o

CD73

oG «a, o

OO

£CDOCCtfGOCQCD

X) OCQ mCD 43

£ »o CD

G

iH C

CD

G

a> ftCD ^CQ ra

CQ g

CD ^

!H Crf

3 CQO Ctf

rt G3 cd

§3 §..H CD ii•^ T3 Op< -t-i QJ^ o .£O G CDCQ -iH *O

? S aCQ G

H M§2•s c °

"2 CD 0)'3 -g N

S .rt

Q, O

ttH G

O _OCQ CQ

'§ g CD•M -in O

I «•!•r* -i-i CQ

CD1 rrl

<l> TO

18

CD

21

Page 27: study of resonance light scattering for remote optical probing

This condition, when coupled with that of no overlap, suggests that scatteringwill be observed when the separation from resonance is much larger than thecollision broadening width of the absorptionline. Here also, we must considerDoppler broadening, and in addition, we must consider the absorption lineshifts introduced by the fields of interacting molecules. The statisticaldistribution of these fields introduces a so-called inhomogeneous broadeningwhich has the effect of shifting some molecules into resonance. Wheneverthe contribution of these molecules to the re-emission is significant, it willdisplay characteristics of fluorescence. This type of broadening is oftencalled "statistical broadening" (ref. 12). It will be discussed in somewhatmore detail subsequently. However, we note that statistical broadening doesnot necessarily fall off as fast as Doppler broadening in the line wings. Henceits effect can extend much farther from resonance. The detailed theory ofstatistical line broadening in the far wings of a line is difficult, and so far in-complete, prohibiting definite conclusions about the transition from scatteringto fluorescence. On the other hand, our arguments suggest that experimentalobservations of this transition might provide valuable information about thebroadening processes.

Finally, at the third level of conjecture, we consider excitation by mono-chromatic (and hence continuous, rather than pulsed) radiation. In this case,the distinction between fluorescence and scattering <ian be made only on thebasis of sensitivity to collision effects (within the criteria adopted in thisreport). The classical analysis developed previously bears no obviousapplication to this case. Indeed, in order to relate the quenching of mono-chromatic excitation to a temporal behavior of re-emission, we would haveto consider individual absorption events, which are distinctly quantum-mechanical in nature. Consequently, we believe that the transition fromscattering to fluorescence in the case of monochromatic excitation is properlytreated only by a quantum-mechanical analysis of intermolecular effects inthe far wings of an absorption line. Such an analysis is not yet available inthe literature, and is regarded as beyond the scope of this investigation. How-ever, in its absence it is tempting to speculate that quenching of monochromaticradiation will not differ substantially from that of broad band radiation centeredat the same wavelength, provided there is no significant overlap of the latterwith the absorption line.

If this and previous speculations are correct, the transition fromscattering to fluorescence should occur at a separation from resonanceseveral times larger than the absorption line width, the emission linewidthor the range of significant inhomogeneous broadening, whichever is larger.However, we do not consider this general conclusion to be firmly established.Preliminary indications from a different, quantum analysis suggest that itshould be regarded with healthy skepticism until tested by additionaltheoretical analysis and experiments.

22

Page 28: study of resonance light scattering for remote optical probing

Simple Quantum Mechanical Calculations

In this section we use relatively simple quantum-mechanical results forscattering and fluorescence to draw some useful conclusions about thesephenomena. The basic quantum-mechanical expression for a differentiallight scattering cross section can be written in the form

I (D'>rn . (D>>fr <D*>rnw rn

(38)

Here DI and D2 are components of the dipole moment vector in the directionsof unit vectors ei.and e2 which designate the directions of polarization ofincident and scattered light, respectively; e. g.

D! = D • G!

The symbol (Di)rn, for example, designates a matrix element of DI betweenthe initial state n and an intermediate state r, and uu = u)r - u)n, where ftuuris the energy of the state r. The quantity Yr is the line width for the state r.To correspond to a practical case this cross section must be averaged overthe distribution of initial states (n) of molecules (atoms) under observationand, for each initial state, it must be summed over all final states (f) whichcorrespond to scattering in the spectral range of interest.

The cross section is defined such that the intensity (watts/steradian) oflight scattered per molecule in corresponding transitions and with appropriatepolarizations is given by

J»2 = a12Ii (39)

where Ii is the. incident irradiance (watts/m2).

Simplification near resonance. - Sufficiently near resonance, one or afew terms in the sum over r in Eq. (38) will often predominate. Unless thegas is highly excited, these terms must be of the first type; i. e. , withdenominators of the form u)rn - uui + iYr/2. In such cases, if only one inter-mediate state is involved for each initial-final state pair, we may write theeross section in the form

(D2)fr rn

23

Page 29: study of resonance light scattering for remote optical probing

In particular, the total cross section, that is, the cross section integratedover all scattering angles, and summed over scattered light polarizations, isgiven by

STT Dfr2

Drn

2

9 c 4 f t 2 (u)rn-u))2 + Yr/4(4.1)

where |Drf |2 designates the absolute square of the vector matrix element;

i. e. ,

Drf = D ,• D- .rf rf <Dk'rf

k=x, y, z

(42)

Note that the absorption oscillator strength for the transition n-»r is definedbv

2m a)rn

rn 3 e ft Drn

(43)

Thus we obtain for the total cross section

2rr 0)2arf Yr/4~|

f f ,rn rf (44)

In this derivation we have excluded consideration of degeneracy arising fromvarious orientations of the total angular momentum of the molecule. For-tunately, it is possible to include this ubiquitous type of degeneracy rigorously.Averaging over initial orientations (magnetic quantum numbers) and summingover final orientations, we find that the result is a^ as given by Eq. (44),multiplied by gf/g , where the g's are the state orientation degeneraciesassociated with the total angular momentum J; e. g.

' g = 2J + 16r r

The near resonance approximations uurn

write Eq. (44) in the formand oirf = 0)3 can be used to

= 2TT — I =*-mc Jrn/

f f ,rn rf(45)

24

Page 30: study of resonance light scattering for remote optical probing

where we have neglected the linewidth yr- It is interesting to note that thisdiffers from the classical result of Eq. (28) only by the factor

3 _L (22- f fe V tUi / rn rf

which is near unity for a strong transition.

(46)

Comparison of Expressions for Scattering and Fluorescence. - UsingEq. (41), we can write an expression for the rate RSC at which photons arescattered per molecule in an incident beam of irradiance 1^ (watts/m2).Thus

Z Tan- (47)

where u)2 = uui - Wfi and the sum over f extends over all final states for whichi«2 > 0. Substituting from Eq. (4 1) for the cross section, we obtain

R = Ii8TT

sc 1 9 c 4 f t 3 (uarn-u)1)2+Y2

r/4) (u>i-u>fi D

fr(48)

The radiation linewidth y is given by

Dfr

This result and the near resonance approximation

) ^ (a - uurf fi

can be substituted into Eq. (48) to obtain

2 v /2trB

3 eftDrn

(49)

(51)

But this expression is equal to that for the rate at which photons are absorbed,given by first order perturbation theory, and therefore it is also equal to therate of Rjp of fluorescence in the absence of collision effects. This resultsuggests that Eq. (38) describes both scattering and fluorescence in the

25

Page 31: study of resonance light scattering for remote optical probing

absence of collisions. In this sense, scattering and fluorescence can beregarded as the same phenomenon in different behavior regimes, as alreadysuggested by the classical analysis. However, in actual experimentalsituations, inhomogeneous broadening and multiplicity of intermediate levelscan lead to a variety of conditions, such that both fluorescence and scatter-ing can appear simultaneously upon excitation near resonance. Furthermore,we must consider processes in which re-radiation excited in an absorptionregion is terminated quickly by an internal process such as dissociation. Insuch cases this radiation can display the properties of scattering, eventhough excited "on resonance. "

Ratio of Scattering Excited Near Resonance to (Quenched) FluorescenceExcited on Resonance. - An estimate of this ratio is useful for experimentalpurposes; for example, one must consider the fluorescence excited by lightin the wings of the spectral distribution of quasi-monochromatic incidentlight in comparison to scattering excited on line center. This estimate alsosets the stage for later considerations of excitation within the wings of anabsorption line. We develop this estimate in terms of the integrated absorp-tion coefficient K, defined as follows. If the transmission of a monochromaticbeam of light through length L of a gas is given by

T = e-k(

where p is the pressure in atmospheres (referred to STP), then (ref. 12)

( / 2 \

K = /k(uu)du) - 2ll2c(^r) NoFrfrn (52)

Here the integral extends over the line, NO is the number density of moleculesat STP and F'r is the fraction of those molecules in the initial state r. Anapproximate value for k(uo) at its maximum can be obtained by dividing thisintegral by the total linewidth T (including collision broadening). This maximumvalue can be equated to N0Fn(aa) peak, where (aa) peak is effective the absorp-tion cross section at line center. Thus we obtain

( „ ) ' . ~ L/J^Y, (53)a peak T \ me J rn

The probability that the absorbed light will be re-emitted in a transitionr-»f (before quenching) is

PE~ W" (54)

providing quenching is the dominant decay mechanism (vc$» yr)- Here yrfis the transition probability per unit time for a radiating transition r-»f, yr

26

Page 32: study of resonance light scattering for remote optical probing

is the total radiation probability per unit time, v is the collision frequencyand § is the probability of quenching per collisioa Note

2(X)2rf frf / e2 \ ....Yrf ~ f — T ) (55.)

Therefore

E cv $ 'c

The total cross section for fluorescence re-emission in the transition r-»fis then

4-n2 cu2 „ / 2 \2

f peak apeak E TV I- \ mcz / rn r^

Assuming that the line broadening is dominated by quenching, we set

Then8nV / e2 x2

<°f>peak - -J3-T ^j fra frf (58)

The ratio of the cross section for re -emission excited off resonance, givenby Eq. (45), to the peak re-emission as given by Eq. (58) is

(of'peak

where AID is the separation from resonance. To recapitulate, the followingassumptions were made in deriving Eq. (59):

(a) Quenching is the dominant de -excitation mechanism.

(b) Simple Lorentz collision broadening model is applicable.

(c) Collisions are the dominant broadening mechanism.

If we assume, moreover, that NRRS is observed at a separation fromresonance equal to 10 times the linewidth F, and I =• 1, then

27

Page 33: study of resonance light scattering for remote optical probing

io-3 ( G O ),peak

Actually, the effects of inhomogeneous broadening may reach out to a furtherseparation than 10 T, in which case Eq. (60) is a high estimate. Neverthelessit illustrates that where a discrete strong absorption line is involved, suchthat the (quenched) fluorescence is likewise strong, there is a good chancethat NRRS will be observed.

Near Resonance Raman Scattering From Atoms

Previously, we have used a slightly more detailed version of Eq. (45),which includes the angle and polarization dependence of the scattering, tocalculate the electronic Raman scattering cross sections for the followinggroup of atoms: boron, aluminum, gallium, indium and thallium. Thisseries is characterized by low-lying excited electronic states and accessiblehigher states through which RS can proceed. These properties are evidentfrom the level diagrams shown in Fig. 7. The NSSR cross sections calculatedfor aluminum atoms (ref. 18) which are shown in Fig. 8 , are typical of thegroup. It is significant that an enhancement of about IO8 over N2 vibrationalRS is predicted at separations from resonance on the order of 1 nm. Thecorresponding Raman shift near resonance is about 2 nm. Although an ex-periment appears feasible to verify this strong re-emission, and to deter-mine whether or not its characteristics are those of scattering, none has yetbeen reported to our knowledge.

Near Resonance Raman Scattering From Molecules

In the situation discussed above, the oscillator strengths are large (~1)and the absorption lines distinct and well separated. For molecules thesituation for this type of NRRS is generally less favorable. Typically,molecule absorption lines are closely spaced, individual line oscillatorstrengths are much smaller and, as a rule, only a moderate fraction ofmolecules are in initial states which contribute to a particular absorptionline. Thus we are led to consider NRRS in proximity to an absorption band,rather than to a line. Typical bands of interest for the molecules O3, SO2,and NO (refs. 19, 20, 21) are shown in Fig. 9 a, b and c.

There is a complication in the calculation of NRRS from a band, becausewithin the absolute square in Eq. (38), some of the terms in the sum overintermediate states (r) may be positive and others negative, producing inter-ference effects. (In the general case, these terms can be complex withvarious phases. ) These effects are discussed in somewhat more detail inAppendix B. However, at a separation from resonance which is large

28

Page 34: study of resonance light scattering for remote optical probing

Go In

I "*

CM

51O

CM

£• o

CM

* t

4d7i

'* tor

01

6d2

I '/2

0 1/2

Figure 7 Level diagrams for aluminum, gallium, indium and thallium,showing transitions which lead to shifted re-emission (Raman scatter-ing or fluorescence).

29

Page 35: study of resonance light scattering for remote optical probing

3800 3900 4000 4100WAVELENGTH OF INCIDENT LIGHT IN ANGSTROMS

Figure 8 Re-emission cross sections for aluminum calculated from oscil-lator strengths. The cross sections azz corresponds to re-emission withpolarization parallel to that of incident light, whereas azx corresponds tore-emission with polarization perpendicular to that of incident light.

compared to the width of the nearest band, but small compared to the separa-tion between bands, it can be shown that these interference effects arenegligible, such that the contribution of the band to scattering is proportionalto its overall strength. Thus, for a first approximation, we estimate NRRScross sections on this basis. Assuming that the initial and final electronicstates are the same, and neglecting rotational structure, we write Eq. (45) inthe form

. 2P l < r | n > |n (61)

n, r

Here f is the electronic absorption oscillator strength, Pn is the fractionaloccupation of the vibrational state n, and | <r | f ) |? for example, is the overlapintegral between vibrational states r and f. The effective resonance frequencyis denoted simply by UJR. The overlap integrals are normalized such that

<a|b> I2 = 1

30

Page 36: study of resonance light scattering for remote optical probing

30

u

* 20

8 "0xo

nl , I . I . I . I i I i220 240 260 280 300 320 340

(0) WAVELENGTH IN NM

220 240 260 280

(b) WAVELENGTH IN NM

300

WAVELENGTH IN NM

225 220 215 210 205 200 195 190 185 180 175 170

o

</> r (o,o)BAND

( C )

Figure 9 Ultraviolet absorption bands of (a)SO2, (b) Q, and (c) NO,partially derived from information in refs. 19, 20 and 21.

31

Page 37: study of resonance light scattering for remote optical probing

Summing Eq. (61) over all final states f, we find that the sum over r canthen be completed, and the cross section for Rayleigh plus vibrationalRaman scattering becomes

rp / \3 / \ 2 I\J \3/ VI \2

aT = 2nPM -^ 1 f 2 =2n *i UV f- (62)

where the second equality has been expressed in terms of wave numbers (e. g.,Vj = w1/2

TTc). The integrated absorption coefficient for the band (here integrated'over wave numbers v) is

" N 0 f (63)

where again NO is the molecule number density at STP. Thus we obtain

°T = -^-2 (~) (Vl-Vv ) K2. (64)

We can write the integrated absorption coefficient in the form

K = <k> 6v

where (k) is the average monochromatic absorption coefficient and 6v is thebandwidth. Then

(65)N;Previously, we assumed that near-resonance scattering would be observed ata separation from resonance equal to ten times an absorption line width.Since a typical bandwidth will be much greater than the width of a single line,we assume here that scattering will be observed at a separation fromresonance equal to the bandwidth. Then

0T = -V 4 <k> 2 (66)TTNJJ ^ .

Now, noting that N0 = 2.69 xlO19 molecules/cm3 at STP and that vt and V2are approximately 3 x 104 cm"1 for incident wavelengths near 300 nm, weobtain

oT(cm2) - 10-30 <k> 2 (67)

for this near UV spectral region, where k is in units of cm-1 atm"1 to base eand referred to STP. Equation (67) can be used along with the results for SO2

32

Page 38: study of resonance light scattering for remote optical probing

and O3 shown in Fig. 9 to obtain an estimate of the enhancement, if any, ofthe scattering from these molecules. Of course, in general the effects oflimited resolution (spectrometer pass band wider than individual absorptionlinewidths) must be carefully considered in determining (k) from absorptionmeasurements. However, the absorption bands for SOa and O3 in Fig. 9 havebeen shown to be quasicontinuous by high resolution measurements. Thuswe can take values of <k) directly from these figures. For SO2 we obtain

(k> =- 20 cm"1 atm"1

and a T =-4 x 10"28 cm2.

But the value of OT for Rayleigh plus Raman scattering from SO^ extrapolatedfrom measurements in the visible is several orders of magnitude larger. Thusthis estimate suggests that there will be no significant enhancement of thescattering from SC>2 under the near-resonance conditions assumed for thisexample.

For O3<k> =-200 cm'1 atm"1 .

In this case Eq. (67) provides the estimate

aT =- 4 xlO'26 cm2

which is approximately equal to the extrapolated cross section. Thus littleenhancement of the total scattering is predicted, although Raman scatteringmay be moderately enhanced if the Raman-to-Rayleigh ratio increases.

A NRRS cross section can be estimated similarly for NO, using in-tegrated absorption coefficients measured by Bethke (ref. 21). For thelongest wavelength absorption band, ^(0,0), we obtain

<k> - 30 cm'1 atm"1

andaT - 10-27 cm2

33

Page 39: study of resonance light scattering for remote optical probing

This value is about two orders of magnitude smaller than the extrapolatedtotal cross section for NO at 226.5 run (1.6 x 10~25 cm2). Therefore in thiscase there appears to be little hope of useful enhancement.

Alternatively, Eq. (45) can be used to estimate the NRRS for incidentlight at, say, 250 nm. In developing this estimate, we simply sum the con-tributions from the various bands listed by Bethke algebraically, neglectinginterference effects which might make the actual cross section substantiallylarger (or smaller) than estimated. We obtain

0T ~ 3.6 xlO'28 cm2

which is orders of magnitude less than the extrapolated cross section.

Thus the cases considered here are unlikely to produce strongly en-hanced scattering. However, in later sections of this report we will pointout more promising situations.

Approach to Resonance - Excitation within the Broadened Line. - Whenthe exciting wavelength is tuned toward the center of a pressure broadenedabsorption line, at some point interactions between gas molecules will beginto have a significant affect on the re-emitted light. In particular, there arespectral regions where pressure brbadening and quenching have comparableopposite effects on the intensity of this light, such that the intensity permolecule can be insensitive to pressure over a broad pressure range.

This phenomenon can be explained within the context of the Lorentzline-broadening theory, following an analysis similar to those developed inrefs. 11 and 22. The absorption cross section in the presence of Lorentzbroadening is given by

a =•Tl* 'cS.rn

2F/2TT . ; ' . . ' .

(u)i-tt)rn)a + T2/4

(68.)

Here Yrn *s tne probability per unit time for the radiating transition r-»n,and T is the total width of the upper level r. The effective cross section forlight re-emission associated with the molecular transition r-»f is given by theproduct of the absorption cross section and the probability PL that thetransition r-»f will occur before some other event (e. g. collision or radiation)alters the state of excitation; thus

34

Page 40: study of resonance light scattering for remote optical probing

(69)

Here the subscript L indicates re-emission into a particular line.

It is possible to derive a simple expression for PL if one assumes thatthe various decay rates associated with the excited state r are independentof the excitation frequency. The probability per unit time of decay by radia-tion is Y- Non-collisional decay by processes such as dissociation and pre-dissociation can occur in some cases; the probability rate for these processeswill be designated by F^. Collisions induce decay by quenching, at aprobability rate FQ. Elastic collisions make a contribution FE to line broaden-ing, and other collisions produce small changes in the excitation energy ata rate F-g, typically creating a broad band of fluorescence upon which thevarious lines are superimposed. Thus the total linewidth is given by

r = Y + r + r + r + r

If the rate of excitation is R, then the steady state population Nr of moleculesin state r is given by

R = Y -N + (Y-Y *)N + r N + F^N + F_N = (r-r ) N' r f r r f r D r B r Q r e r

Here we assume that repopulation of state r by collisions is negligible.Then the probability of decay by re-emission r-»f is given by the ratio ofthis rate to all decay rates; i. e. ,

frfr-r. (70)

E

Likewise, if we assume that the various decay rates for all excited statesare equal, the probability of decay by re-emission into the broad continuumfollowing excitation of state r is

B

From Eqs. (68), (69) and (70), we obtain for the line fluorescence

Yrn F/2TT rf(71)

35

Page 41: study of resonance light scattering for remote optical probing

This result leads to interesting predictions, for if collision broadening andquenching predominate, then to a good approximation

r = Kpwhere p is pressure and e, q and K are appropriate proportionality constants.Substituting the first and last of these results into Eq. (71) we obtain

1

j

Y urn " ' " (72)

L>i-«>rn)2 + K2p2/4.This equation predicts that for excitation well out on the line wings, whereuj-uJrn »Kp, the resulting line fluorescence per molecule will be independentof pressure. Furthermore, to the extent that the last term approaches unity(insignificant proportion of elastic collisions) it will be independent of gasconstituency. A similar result obtains for the broad continuum fluorescence.These results are not changed by inclusion of Doppler broadening providedthat the separation from resonance is also large relative to the Doppler width.

Although this particular result has not been emphasized previously to ourknowledge, the basic idea is not new. Mitchell and Zemansky (ref. 12) discussthe interplay of line broadening and quenching in qualitative detail, and the re-sults of Fouche, Herzenberg and Chang (ref. 11) are very similar to those pre-sented above. However Fouche et al. did not include the elastic collisionswhich, within this simple theoretical model, are necessary to account forinitial increases in the fluorescence cross section with background gas pressureas observed by St. Peter' s et al. (refs. 14, 22).

The simple Lorentz broadening theory is still utilized as a useful firstapproximation close to resonance. Thus, it is interesting to compare Eq.(72) with the quantum mechanical result for scattering, which should describethe same effect far from resonance. If transitions through a single inter-mediate level dominate the scattering (as in the case of Al atoms describedpreviously, even tens of nm from resonance) then the scattering crosssection is, from Eq. (45),

(u)i - ujrn)

Using the relationship between oscillator strength and radiation line width.2

Y = — MV f (74)rn c V r n ^ 2 / rn

36

Page 42: study of resonance light scattering for remote optical probing

and the equivalent expression for yrf» we obtain

TT2C2Y U>2 Y ,/2TTrn rro = 3 — < 7 5 >

o d\ f r i) — ni f1

This result differs significantly from Eq. (72) by the absence of the factorK/(K-O- The difference reflects a failure in one or perhaps both-theoreticalanalyses which must be examined by a more detailed theory.

Effect of Statistical Broadening. - Up to this point, we have used theLorentz broadening theory to describe the effects of interactions betweenmolecules on fluorescence. There is a complementary approach to linebroadening which leads to somewhat different conclusions. In the rudimentaryform of this theory, the relative motions of molecules are neglected. Thefields at a molecule produced by its near neighbors introduce relative shiftsin energy levels such that the absorption line for that molecule is also shifted.The overall broadened line is produced by the superposition of these shiftsweighted by the statistical probability of the configurations producing eachshift. This approach has been called the statistical broadening theory(ref. 12).

Thus, if we designate the coordinates describing a particular configura-tion of neighbors around a molecule by X, the probability of this configurationby P(X) and the resulting shift by 6(X), the absorption line is given in thesimplest formulation by

o (ID) =arn Y./2TTrdXP(X) ;— (76)

[ u)-i»rn-6(X)]2+ Yr/4

Except for the contribution of the natural line width, this type of broadeningis of the type often called inhomogeneous broadening; that is, the absorptionline of any particular molecule at a particular time is much narrower thanthe composite line width. Doppler broadening, which is also inhomogeneousin nature, can be incorporated within this formulation by extending thevariables indicated by X to include the velocity of the absorbing molecule,and including within 6(X) the corresponding Doppler shift.

37

Page 43: study of resonance light scattering for remote optical probing

Equation (76) differs from its Lorentz broadening counterpart in thatit does not predict increasing absorption with pressure far from resonance.Furthermore, it can account for the absorption line asymmetry and shiftwhich are observed in many pressure broadening experiments. However itrepresents an extreme case (static condition) just as does simple collisionbroadening (instantaneous collisions). Many physical situations can be ex-pected to fall somewhere in between. Such cases can be represented con-ceptually by adding a term to yr which depends on the rate of change of theconfiguration X.

It is the task of detailed line-broadening theory to predict the behaviorof the inhomogeneous and homogeneous contributions to the line shape. Thegoal of our theoretical considerations up to this point has been to use simple,general results to illustrate the relationships between scattering andfluorescence, as these terms are defined herein. We have also tried tomake clear the intimate relationship between fluorescence excited nearresonance by narrow wavelength sources, and the basic characteristics ofline broadening. The continuing development of tunable lasers is opening thisfield of research into what should be an exciting and informative subject.

Excitation into Dissociative Continuum. - When a molecule is excitedinto a dissociative state by absorption of light, it must either re-emit thisenergy quickly or fly apart, converting an appreciable fraction of its energyto kinetic energy. The time required to dissociate is on the order of 10"11

to 10~12 seconds. Thus, in this situation there is little time for collisions toquench or alter the re-emission, with the result that it assumes thecharacteristics of scattering.

Holzer, Murphy and Bernstein (ref. 4) have identified and examined thissituation in re-emission from halogen vapors 12, Br2, C12, IBr, IC1 andBrCl. The scattering is characterized by strong overtones out to manyharmonics, introduced by transitions through the continuum of outgoingspherical waves which characterize the dissociative states. The enhancementof this kind of scattering can be moderately strong. Holzer, Murphy andBernstein found that when I2 is excited by 488.0 nm radiation, the crosssection for the fundamental vibrational RS is about 4.4 x 10~28 cm2/sr, 100times larger than that for the vx band of methane, or about 800 timesstronger than that for N2 at that wavelength. However, cross sections forthis type of scattering from Br2 and C12 are much smaller.

Other molecules which are conveniently excited above their lowestdissociation limits are O3 (~1000 nm) and NO2 (398 nm) (ref. 23). Conflictingresults exist for NO2 excited at 337.1 nm, with unpublished Vi vibrationalcross sections quoted from 1 to 1000 times that of N2. One possible sourceof difficulty with these measurements is that N2O4 exists in equilibrium with

/ 38

Page 44: study of resonance light scattering for remote optical probing

NO2, and the equilibrium constants are sensitive functions of (NO2/N2O4)partial pressures and temperature. Furthermore, N2O4 absorbs morestrongly than NO2 near 337.1 nm (ref. 24).

Finally, the RS cross sections for O3 excited near its very strong dis-sociative continuum centered at 250 nm have not been measured yet. However,in Chap. IV we develop an estimate of this cross section that suggests it mightbe very large.

Scattering following excitation into the dissociative continuum has beenexamined theoretically by a number of authors, with particular attention givento I2. Behringer (ref. 25) developed a theoretical expression for its intensity.Jacon, Be r jot and Bernard (ref. 8) developed a quantitative expression for thescattering intensity as a function of the incident wavelengths, and demon-strated agreement between this expression and their experimental results.In particular, they pointed out the importance of off-resonance (virtual)terms in the scattering amplitude. Finally, Kiefer and Bernstein (ref. 26),and Williams and Rousseau (ref. 27) have explained the vibrational rotationalstructure of the vibrational and overtone " lines. "

Scattering Over Quenched Fluorescence. - Jacon et al. (ref. 9) havepointed out that when fluorescence excited on resonance is strongly quenched,its intensity can be reduced to a point where scattering through nearby off-resonance states is comparable. This behavior can be predicted by writingthe quantum-mechanical expression for the cross section in the form

. , ( f l D l r X r l D l i ) T' (f J D J g ) (glDJi):^ r" ~u3 -liu -ir /2—

61 l S (77)

I 2

+ antiresonant terms )

Here we have summed over the limited group at final states (f) for whichwi - ^f i is in the frequency range of interest. The contribution of the inter-mediate state in resonance (r) is separated out of the sum over intermediatestates, and we have inserted total widths F (including collision broadening)into the resonance denominators in order to take account of Lorentz broaden-ing. Jacon et al. consider Eq. (77) to represent both fluorescence andscattering.

The widths F in Eq. (77) can be expressed in the form

r = Y + KPr r

where Yr is the total width for non-collisional decay.

39

Page 45: study of resonance light scattering for remote optical probing

At low pressure the resonance term will clearly predominate (providing theline separation is much larger than line widths), leading to strong fluorescence.However, at high pressure the resonance term will be much smaller, whereasthe sum over off- resonance terms will change relatively little if u)gi - uj» kpfor all significant contributions. Thus it appears that there are cases in whichnear -resonance scattering can predominate over strongly quenchedfluorescence.

Berjot, Jacon and Bernard (ref. 10)have used this argument to explain thechanges in the re -emission of 1% excited by 501.7 nm radiation as the back-ground pressure of argon is varied from 0 to about~40 atmospheres. Thechanges in depolarization, lineshape and overtone sequence are all character-istic of the transition of the dominant contribution from fluorescence (at lowpressure) to Raman scattering (at high pressure). However, Jacon et al.(ref. 9) leave the impression that the intermediate states which contribute thedominant scattering at high pressure are part of the discrete set, as impliedby Eq. (77). In fact, the polarization, intensity and lineshape variation withpressure all indicate that the dominant scattering contribution arises fromtransitions which begin from excited vibrational states and reach the dis-sociative continuum. Thus, we believe that Jacon et al. have actually observedRS following excitation into the dissociative continuum, rather than NRRSfrom discrete states.

Fluorescence

Dependence of Fluorescence on Gas Pressure. - The basic equationsgoverning the dependence of fluorescence on gas pressure have been givenpreviously. From Eq. (71), the effective cross section per molecule for linefluorescence excited by a monochromatic incident beam is

aT (u)) = A - - . „-£- - (78)L (u^-iWj.) + 1/4

under the assumption of Lorentz broadening. Here A represents factors thatare independent of pressure, and YL represents the probability per unit timefor decay by emission into the observed line. The pressure dependence ofthe contributions to the linewidth T can be expressed by

r = bP (79)

B F

40

Page 46: study of resonance light scattering for remote optical probing

where Y is the total width for non-collisional decay (radiation + dissociation,etc. ) and p is gas pressure. Equation (78) is applicable at line center,providing that collision broadening dominates over Doppler broadening, andin the line wings beyond the range where they are effected by Dopplerbroadening. If, in addition, excited state decay is primarily throughcollision-induced processes, then for excitation on line center (u)!=ujr)

0L oc i/p2 (80)

Likewise, on the wings we find that a^ is independent of pressure, asdiscussed previously.

On the other hand, if the spectral distribution of the incident beam isconstant over the absorption line, the effective "cross section" is obtained byintegrating a^ (uji) over frequency. This operation yields

(81)

It should be noted that OL now has dimensions of area (x) frequency, since itmust be multiplied by incident irradiance per unit frequency. This resultproduces the pressure dependence of the Stern-Volmer equation (ref. 12); .i. e. ,

Y+(b+q)p(82)

Likewise, if the absorption is continuous or quasi-continuous (composed ofmany lines with line separation < line width) such that the absorption permolecule is not pressure-dependent, we obtain

YT0T = A p(uu) rr-^— (83)^ E

where p(uo) is an effective line density. Thus, in this case also the Stern-Volmer pressure dependence is predicted. It should be noted that the resultsfor the last two cases do not depend on lineshape, but rather on the integralover it. In that sense they do not depend on the assumption of Lorentzbroadening. A case of major present interest involves fluorescence from oneof the minor constituents in a carrier gas (air). Under the assumption thatcollision rates are additive,

41

Page 47: study of resonance light scattering for remote optical probing

e2p2 + . . . + ecpc

TB = blPl + b2p2 + . . . + bcpc (84)

TQ = qiPi + q2P2 .+ • • - + qcPc

Here pi, p2 . . . represent partial pressures of minor gases, and pc, the partialpressure of the carrier gas. Likewise GI • bx, qi , etc. , represent correspond-ing coefficients for components of the line width. In cases where the carriergas dominates the quenching, " -----

( q + b ) p(85)

and consequently, the fluorescence per molecule is independent of minorconstituency. When this result is applicable to atmospheric monitoring, itimplies that only a simple and rather small (< 10$) pressure broadeningcorrection must be made to fluorescence data in determining the moleculenumber density of a minor constituent. This is the favorable situationfor probing applications which was pointed out previously.

Depolarization of Scattering and Fluorescence. - The polarizationcharacteristics of scattering and fluorescence are relevant to present interestfor two reasons. First, polarization can often be used to identify what typeof process is being observed; second, processes can be separated from eachother and/ or background on the basis of polarization. For example, itshould be possible to view nearly unpolarized fluorescence from NO2 overhighly polarized Raman scattering from CO2, even though both occur in thesame spectral region, by using polarization filters to reject the CO2

scattering.

The polarization of scattered light and fluorescence is partly character-ized by its depolarization factor. There are two alternative definitions ofdepolarization. The one we shall use is in terms of an experimental con-figuration involving a linearly polarized incident light beam and scatteredlight propagating in a direction perpendicular to the polarization andpropagation direction of the incident beam. The depolarization p is then thefollowing ratio: the intensity of that scattered light with polarization perpen-dicular to the incident beam polarization, divided by the intensity of thatscattered light with polarization parallel to the incident beam polarization.This definition is illustrated in Fig. 10 . In terms of the symbols used in thatfigure

P - IJl (86)

42

Page 48: study of resonance light scattering for remote optical probing

OBSERVED SCATTEREDBEAM POLARIZATIONS

INCIDENTBEAM

POLARIZATION

ySCATTERED

BEAM DIRECTION

INCIDENT x

BEAM DIRECTION

Figure 10 Scattering geometry for definition of depolariza-tion.

The definition of the alternative depolarization p1 is similar except that theincident beam is unpolarized. The relationship between the two quantities is

(87)

The significance of these depolarization factors extends beyond theexperimental configuration in which they are defined. In fact, the de- "•"'""polarization is a fundamental characteristic of the scattering, or fluorescence,determining for example its angular dependence (ref. 18). The depolarization(p) for vibrational Raman scattering away from resonance is quite small formany molecules, with occasional exceptions where it reaches 3/4. On theother hand, the depolarization of non-resonance rotational Raman scatteringis usually 3/4. "M

The studies of depolarization in scattering and fluorescence trace backto the original theoretical works of Placzek (ref. 28) and Mrozowski (ref. 29)and the early experimental work of Wood and subsequent work ofPringsheim (ref. 30).

The depolarization of scattering and fluorescence from symmetric topmolecules excited near and in resonance has been examined theoretically

43'

Page 49: study of resonance light scattering for remote optical probing

by Seth D. Silverstein. This examination has revealed an effect not previouslypredicted to our knowledge: the dimunition of depolarization in scatteringupon excitation in a dissociative continuum.. This result has been demon-strated in experiments conducted by R. L. St. Peters and Seth D. Silverstein.Their results have been published (see ref. 31 and Appendix D). Below wepresent a qualitative description of their work.

Two types of processes are considered. In the first, the transition isfrom a discrete initial state through a discrete intermediate state to adiscrete final state (DDD transition). In the second, the intermediate stateis within the dissociative continuum (BCD transition). When the excitation is :on resonance, DDD processes lead to fluorescence, while in the case ofexcitation well off resonance they may lead to scattering, as we have defined it.(But see Appendix A. ) On the other hand, the DCD processes lead toscattering of the type identified by Holzer, Murphy and Bernstein (ref. 4).Many of the depolarization results for the DDD case have appeared previously,but in obscure references; these have been re-derived in the present work.

In the DDD case, the depolarization contributed by individual transitionsis independent of the quantum number K which designates the angularmomentum component of the molecule parallel to the figure axis of thesymmetric top. For K ^ 0 there are in general three Q branch transitions,which we denote by Qa (a = ±1, 0). These components correspond to the totalangular momentum sequences shown below.

+ 1\

The individual depolarization contributions for these Q-branch componentsdepend on the initial total angular momentum quantum number J. Fora = ±1, both go to 3/4 asymptotically at large J, while the a= 0 componentgoes to 1/3. For K = 0 in both electronic states there is no Q-branch ab-sorption and the component corresponding to a = 0 is absent.

Besides the Q branch transitions, there are those in which the totalangular momentum changes from initial to final state. The entire manifoldof transitions is shown below.

44

Page 50: study of resonance light scattering for remote optical probing

J + 2

,J + 1 ^ J + 1

-» J-

'J - 1^ "^J - 1

J - 2

This manifold obviously results from the emission-absorption selectionrules AJ ± 1, 0. I fK=0, in both upper and lower electronic states, then theselection rules reduce to AJ = ±1, and the J1 = J±l transitions vanish, alongwith the J-»J-»J transition. The depolarization of AJ = ±2 transitions, andthe asymptotic depolarization of the remaining Q-branch transitions is 3/4.

Since the Q-branch components start from the same initial state, and endon the same final state, their contributions can interfere within the absolutesquare of the scattering equation. (See Appendix B for a general discussionof interference effects. ) However, the resonance situation will pick out onlyone of these transitions in the ODD case, providing that the separation of theupper states is larger than the linewidths involved.

In the case of 1% excited from the ground electronic state by wavelengthslonger than the dissociation limit, the transitions are of the ODD type andK = 0 in both upper and lower electronic states. Furthermore, the rotationalconstant of I2 is very small because of the large nuclear mass of iodine.Thus high angular momentum states are thermally excited. The transitionsobserved in recent experiments on I2 using argon laser excitation are nearlyall high J transitions, such that the asymptotic high J results for Q-branchdepolarizations are applicable. Thus the depolarization of each componentof the doublet which results (corresponding to transitions J-* J+l -* J+2, 0 orJ-»J-1 -»J-2, 0, depending on which state is excited) is 3/4.

This depolarization is observed, in fact, at low pressure. As (I2 orbackground gas) pressure is increased, two effects can change the de-polarization. First, elastic collisions occuring between excitation and re-emission cause it to increase. Second, homogeneous line broadening canenable significant contributions from both (J±l) Q-branch components, inwhich case they can interfere within the absolute square, reducing de-polarization. In I2, the separation of the J±l upper states in transitionswhich have been observed is sufficiently large so that the first effect can beexpected to predominate, at least at moderate pressures. In fact, an in-crease in depolarization is observed in this regime.

45

Page 51: study of resonance light scattering for remote optical probing

In the DCD case, the situation is markedly different, because thecontinuum states are effectively degenerate in J. As a result, the Q-branchtransitions through J+l and J-l intermediate states interfere, reducing thedepolarization of the Q-branch to a limiting value of 1/8. The experiments

i we have conducted to observe the Q branch in DCD transitions yield a de-polarization of 0.12 ± 0. 01, in excellent agreement with the theoreticalresult. These experiments necessarily require very narrow (<2cm 1) slitbecause the J1 = J+2 and J' = J-2 branches are shifted only slightly from theQ-branch. Theoretical calculations indicate that the composite depolarization,of all three branches is 1/3 in the limit of large J. This result has alsobeen confirmed by pur experiments with wide slit functions, and has beenobserved by others. Our interpretation of the high pressure 1% experimentsdescribed by Berjot et al. (ref. 10) is that the DDD fluorescence that pre-dominates at low pressure is quenched to the point where DCD scattering,which is not quenched, predominates at high pressure. This interpretationis consistent with their relatively wide slit function observations of thechange in depolarization from 3/4 to 1/3 at high background pressures andtheir observation that higher pressures are required to produce the effect at514. 5 nm than at 501. 7 nm. The latter observation is reasonable becauseDCD transitions are weaker at the larger wavelength because fewer moleculesare available in the higher vibrational states which produce them.

46

Page 52: study of resonance light scattering for remote optical probing

HI - EXPERIMENT

In this chapter we describe experimental results obtained under supportof this contract, and the two separate experimental facilities utilized for thiswork.

Double Monochromator Facility

The Double Monochromator Facility (DMF) has been in use for severalyears, providing measurements of Raman cross sections and scattering line-shapes in flames, among other things. A block diagram of its present con-figuration is shown in Fig. 11.

Light sources. - The primary light sources used with this facility areseparate argon and krypton ion lasers (Coherent Radiation Laboratory Model52) run from a common power supply. The ion lasers feature servo powerstabilization to much better than 1$, and they can be operated in single long-itudinal mode (SLM) using a tilted etalon. This technique provides lines withspectral width and long term stability within 0. 0001 nm, in conjunction withtuning ranges on the order of 0. 01 nm for the stronger lines. The availablespectral lines, corresponding power levels and SLM performance are shownin Table I.

Double Monochromator. - A Spex 3/4 meter double monochromator,model 1400-11, is used to analyze the light diverted from the incidentbeam by gas molecules. This monochromator is fitted with 1200 -t/mmBausch and Lomb gratings blazed for 500 nm (first order), providing a linearreciprocal dispersion of about 0. 55 nm/mm at the exit slit. A periscopicmicroscope focussed on the entrance slits from inside the monochromatorfacilitates alignment, since the image of light diverted from the laser beamis likewise focussed on these slits when optimum alignment is achieved, andthus the slits and scattering volume are viewed simultaneously in focus throughthe microscope. The wavelength setting is sensed directly from the leadscrew by a digital encoder providing 0. 001 run resolution. Scattered light whichhas passed through the spectrometer is detected by a selected RCA C31000Eextended red response photomultiplier in a dewar cooled to -70°C by N2 gasbubbled through liquid N2- The dark count of this tube at room temperatureis approximately 20, 000 cts/sec, but at -70°C it is reduced to about 10 counts/sec. Careful measurements have revealed that the tube sensitivity to visiblelight remains nearly constant over this range. This photomultiplier has asemiconductor first dynode whose high secondary electron production producesa discrete pulse height spectrum. This characteristic facilitates pulse count-ing signal processing.

47

Page 53: study of resonance light scattering for remote optical probing

BLACK SCREEN

/ / REFERENCE LAMP

BUCK BOXLAMP ENCLOSURE

POLARIZATIONROTATION

LASER

DOUBLEMONO-

CHROMATORCOLLECTIONLENS

PHOTOMUL-TIPLIER

CALIBRATIONSCREEN

SCATTERING CELL. THEINCIDENT BEAM IS REFLECTEDVERTICALLY THROUGH THISCELL FROM BELOW ANDTERMINATES IN A POWERMONITOR ABOVE CELL

Figure 11 Block diagram of the Double Monochromator Facility(DMF).

48

Page 54: study of resonance light scattering for remote optical probing

TABLE I

Some spectral lines available from the Ar+ and Kr+ CRL Model 52 lasers usedas primary sources for the double monochromator facility. Lines which havebeen operated in longitudinal single mode (LSM) and, in some cases, LSMtuning ranges and peak powers, are shown in the last two columns.

X .air

350.74/356. 42

351. 11/363.79

457.94

476.24

476.49

487. 99

496. 51

501.72

514. 53

520.83 . .

530.87

568. 19

647. 10

Type

Kr

Ar

Ar

Kr

Ar

Ar

Ar

Ar

Ar

Kr

Kr

Kr

Kr

Power LSM

0. 05W

?

0. 5 W

0. 2 W

0.8 W y

2. 1 W ~0.01 nm

0.8 w y0.3 W 0.005 nm

2. 3 W ~Q. 01 nm

o. 2 w y

0. 5 W 0. 01 nm

0.8 W ~0.01 nm

1.2 W

Peak Power

0.7 W

0. 1 W

0.7 W

0.05W

0. 2 W

0. 3 W

49

Page 55: study of resonance light scattering for remote optical probing

Signal processing electronics. - The preamplified photomultiplier pulsesare brought to a single channel pulse height analyzer (Hamner Model N302)which is usually operated in integral mode to reject small pulses characteris-tic of noise. The accepted pulses are sent to a count rate meter which pro-vides an analog signal for an x-y recorder. They are also counted over presettime intervals and printed along with wavelength readings from the spectro-meter digital encoder. A punched paper tape representation of this printoutis also provided in order to communicate data to a computer for subsequentmanipulation and display.

VacuurrTarid gasr handling system. "-~ The^purpose~of this system is~to pro-vide gases and gas mixtures for observation at the spectrometer. The systemcan be connected through a stainless steel bellows to a mounted scattering cell.In order to withstand corrosive gases, the system is constructed ofstainless steel tubing and Hoke stainless steel valves, with stainless steelSwagelock fittings. Rough indication of gas pressure is provided by thermo-couple gauges, and accurate pressure readings are obtained using a combina-tion of Hg and silicone (Dow Corning704) oil manometers. The system isevacuated by a Welch Duo Seal roughing pump and cold trap. It is necessaryto provide a bypass on the cold trap in order to avoid condensing gases suchas NOg and SC>2 in this trap. After long evacuation, the entire system will holdat < 10 mTorr for days with the pump and cold traps valved off, indicatingthat there are no significant leaks. '

Calibration. - Absolute and relative spectral calibrations of the spec-trometer-photomultipler system are obtained using GE 200W quartz-halogenprojection lamps. These lamps, which are recommended by the Bureau ofStandards as intense and stable secondary standards (ref. 32), were calibratedby Eppley Laboratory, Inc. , and Optronic Laboratories, Inc. (one each).The absolute calibrations were checked by subsitution and found to be consistentwithin 10$. The lamps are mounted in a large (75 cm cube) black box with asmall hole, through which one of them illuminates a diffuse scattering screenat a distance of several meters, as shown in Fig. 11. The scattering screenwas constructed by flowing Eastman White Reflectance Paint (a suspension ofBaSO4 crystals) (ref. 33) onto a 1/2" thick aluminum plate. The area to becovered (9" square) was milled out to a depth of 1/16" and sandblasted. Thebarium sulphate coating was built up in several steps, with intervening slow(~ 1 day) drying in an enclosed space, and finally scraped with a sharped-edgedglass plate back to a thickness of 1/16". Through a direct comparison ofincident and scattered light, this screen was found to approximate a perfectLambertian scatterer to within experimental error of 3$ at a scattering angleof 30°, with incident light normal to the screen, over the spectral range from450 nm to 500 nm, and separately, over the range from 500 nm to 650 nm.These spectral ranges were isolated using filters.

50

Page 56: study of resonance light scattering for remote optical probing

The light from the reference lamp is diffusely scattered by the screenand then passes through the scattering cell. Part of this light is focussedby the collection lens through the monochromator slit. A polarization filternear the lens (on the slit side) is inserted and rotated to determine the relativepolarization response of the spectrometer. Sensitivity to second order dif-fraction (usually negligible) is evaluated by using a filter substitution technique.

This calibration technique provides the following important advantages:

1. The reference count rate and scattering/fluorescence count rate canbe made approximately equal with convenient slit settings and reference lamp-to-screen distance. In our work visible light calibrations were obtained withthe reference lamp placed at a distance of 3 to 4 meters from the scatteringscreen, and the slit settings used in reference measurements were typically300 um xO.5 cm (entrance), 3000 ^m (intermediate), and 300 ;ain (exit).These slit widths are sufficiently large to avoid significant diffraction andpolarization effects.

2. All of the light from the reference lamps collected by the lens andsubsequently focussed through the entrance slit of the spectrometer must passthrough the virtual image of the entrance slit within the scattering cell.Since the scattered light emanates from the same region, spatial variationsin system response should not introduce significant calibration errors.

3. The calibration of the system can be checked using the referencelamp during each measurement, even with the cell in place.

Experiments With NC>2

Nitrogen dioxide exhibits strong absorption through most of the visibleand into the uv. At wavelengths longer than the first dissociation limit(~398 nm) the absorption spectrum consists of many discrete lines; howeverthe distribution is so dense that the fine structure can be seen only at verylow NO2 pressures (< 1m Torr).

Laser-excited fluorescence from NO2 has been studied by several groups,among which are: Sakurai and Broida (ref. 34); Sackett and Yardley (refs.35, 36, 37); Fouche, Herzenberg and Chang (ref. 11); and Stevens, Swagel,Wallace and Zare (ref. 38). Sakurai and Broida described the overall featuresof the fluorescence spectrum, measured low pressure absorption coefficientsat a number of laser wavelengths, and identified an isolated absorption linewithin the 488 nm argon laser gain curve. These authors clearly demonstratedthe strong advantage tuned laser fluorescence provides as a technique for un-ravelling complicated molecular spectra.

51

Page 57: study of resonance light scattering for remote optical probing

Sackett and Yardley carried out a detailed study of the time dependenceof NO2 fluorescence following pulsed excitation, as a function of incident lightwavelength. Using a dye laser with narrow line (<0. 1 nm) output in the blue(440-490 nm) they observed non-exponential decay which showed definite vari-ations with excitation wavelength. In particular, for some excitation wave-lengths they observed a weak component with much shorter lifetime (—1. 25 p.sec)(ref. 36). This observation is relevant to the anomaly, as yet incompletelyunderstood, which exists in NO2, SO2, and CS2 between the strength of absorp-tion and the predominant fluorescence lifetime (ref. 39). In NO2 the latter isabout 100 times longer than expected from the accepted theoretical relationshipbetween these two quantities. Sackett and Yardley's results also suggestthe possibility that there may be spectral regions where the fluorescencequenched by air is much stronger than observed so far.

The theoretical contributions of Fouche, Herzenberg and Chang have beenmentioned in Section II. Their experimental results include measurementsof the absolute intensity of the line and continuum fluorescence componentsexcited at several Argon laser wavelengths, and measurements of air quench-ing coefficients. These results will be discussed in more detail subsequently.

Stevens, Swagel, Wallace and Zare report the characteristics of NO2

fluorescence excited by a narrow band (0. 01 nm) pulsed dye laser operatedin the spectral region from 593. 4 to 594. 0 nm. From the spectral, temporaland polarization characteristics of the fluorescence, they were able to drawimportant information regarding the electronic states and rotational transi-tions associated with absorption in this region.

It is possible that NO2 fluorescence can be used to measure its con-centration in the atmosphere. This possiblity holds particular interest be-cause NO2 is an important primary pollutant, it is also a link in atmosphericphotochemical chains producing more complex and more insidious pollutants,and as yet there are no well developed ways to measure it remotely in theatmosphere that appear to be fully satisfactory. The spectral distributionof re-emission irom a mixture of NO2, NO and N2 at a total pressure near oneatmosphere, excited by 488. 0 nm light, is shown in Fig. 12. Fluorescencefrom NO2 contributes three fairly sharp "lines" and a broad continuum. TheRaman scattering lines of NO and N2 also appear, superimposed on the con-tinuum. It is apparent from this figure that even when strongly quenched,NO2 fluorescence possesses detailed structure which can be used to identifyits presence. On the other hand, both the line and continuum fluorescencecan overlay Raman scattering from other gas constituents, contributing anundesirable background. Thus it becomes doubly important to know the strengthof line and continuum components under conditions typical of those observed byatmospheric probing instruments.

52

Page 58: study of resonance light scattering for remote optical probing

800

600

CISSEC

400

200

NO

_L5000 5250A 5500

Figure 12 Spectral distribution of re-emission from trace amounts of NOand NO2 in Ng at a total pressure of 740 torr. Line fluorescence from NC>2and Raman scattering from NO and N2 are superimposed on wide bandfluorescence from

800 mTORR N02 INAIR AT 740 TORR

4880/1 EXCITATION

N02(Z'|

I I521.0 527.0

522.0 528.5

WAVELENGTH IN NM-*-

Figure 13 Typical spectral trace usedto determine line and wide bandfluorescence intensities from NOg inair. The excitation wavelength is488. 0 nm.

550.6

53

Page 59: study of resonance light scattering for remote optical probing

In previous work we measured the intensity of these components relativeto Og Baman scattering in air at pressures near one atmosphere. In thesemeasurements care was taken to avoid errors introduced by the NO2/N2C>4equilibrium, which is quite sensitive to temperature, and by the tendancy ofexcited NO2 to migrate out of the focal volume before re-emitting light. Themeasurements were repeated using the DMF with various slit settings forthis contract. The results of one of these measurements is shown in Fig. 13.The detailed spectral distribution of the Vj line at low pressure (< 1 Torr)is shown in Fig. 14. Measurements of depolarization in the geometry of Fig.10, obtained by rotating a polarization filter in front of the entrance slit,indicate that the depolarization of this line in air near STPls 1.0 ± 10$.Taking into account the spectral sensitivity variation of the spectrometer, therelative line widths and different polarizations, the results of Fig. 13 indicatethat the NO2 fluorescence intensity in the Vi "line" is 110 ± 30 times strongerper molecule than Q-branch vibrational RS from C^. This result applies for488. 0 nm excitation in air at 740 torr, in the right-angle scattering geometryof Fig. 10, In order to obtain an absolute result, we note that the correspond-ing O2 cross section has been measured in our laboratory to be (ref. 5)

6.8 x 10-31 cm2/sr ±10%

This value, for scattered light summed over polarization, is in good agree-ment with recent results obtained at two other laboratories. Thus we con-clude that the effective differential cross section for NO2 line fluorescenceunder the conditions specified above is

7.5 x 10"29 cm2/sr ±35$

or about 130 times enhanced over the N2 vibrational RS. Because of the strongdepolarization of the fluorescence under these conditions, it should be nearlyisotropic. Therefore, the corresponding total cross section should be, to agood approximation, 4n times the differential cross section, or

9.4 x 10~28 cm2

These results are in fairly good agreement with those obtained by Fouche,Herzenberg and Chang (ref. 11). This group measured the air- and self-quenching constants of NO2 line (Vj and v2) and continuum fluorescence, andthe magnitude of these fluorescence components relative to N2 vibrationalRaman scattering at a NO2 pressure of 1 torr. From their results we cal-culate that the effective NO2 differential cross section for the Vj line relativeto the corresponding N2 vibrational Raman cross section (for a trace of NO2

in air at 740 torr and excitation at 488 nm) is approximately 190. The equiv-alent ratio for the V2 line (750 cm'1) is, from their data, 132 and for thecontinuum fluorescence in a Inmbandata shift of 720 cm"1, 120. The agree-ment for absolute line intensities is reasonable but our results indicate a

54

Page 60: study of resonance light scattering for remote optical probing

I I I520 521 522 523 524

WAVELENGTH IN NM525

Figure 14 Detailed spectral distribution of the NOa Vi-line fluorescence ex-cited by 488. 0 nm radiation. The spectral slit width for this measurementwas triangular with 0. 15 nm full width at half maximum.

continuum intensity about 6 times larger than theirs for a 1 nm band at 720cm"1 under the conditions cited above.

As long as air is the primary quenching species, the effective NC>2 crosssections should be inversely proportional to collision frequency with airmolecules. It is interesting to note that at an altitude of 30 km, where thecollision frequency is reduced to about 1$ of its sea level value (~1010 sec"1)the effective cross sections for line re-emission from NC>2 should be en-hanced by about 104 over the N2 vibrational Raman cross section.

Experiments With Iodine Vapor

The vapors of halogens such as I2, Br2, and C12 exhibit absorption in thevisible, arising from 3Ho+ -1S+ electronic transitions. The dissociationlimits for these transitions are in the blue-green spectral region. Upon ex-citation at wavelengths longer than the dissociation limit, fluorescence hasbeen observed, whereas at shorter wavelengths, the re-emission is charac-teristic of scattering (ref. 4).

55

Page 61: study of resonance light scattering for remote optical probing

Iodine and Bromine vapors are convenient subjects for studies of absorp-tion and consequent scattering/fluorescence because they are relatively safeto work with, they are chemically compatible with glass, their vapor pressureis sufficient for many experiments at room temperature, and the distributionof their absorption lines is sufficiently dense so that one and often severalabsorption lines are found within the gain curves of many argon and kryptonlaser lines.

Our experimental studies have involved only I2. Although this moleculeseems to have a somewhat greater density of lines than NO2, individual linesare easily resolved in absorption in I2. For example, the absorption of I2

within the 514. 5 nm gain curve shown in Fig. 15 indicates the presence of fourapparent lines within and on the edge of the gain curve. (Actually, some ofthese lines result from multiple transitions in near coincidence. ) These linesremain resolved at the highest observed I2 vapor pressure, 2. 3 Torr. Thesituation in NO2 is notably different. Sakurai and Broida (ref. 34) observedthe presence of an isolated NC>2 line within the 488. 0 nm gain curve and ob-tained some information about its shape by observing the transmitted lightthrough a Fabry Perot interferometer. We have also observed this line usingthe method of tuned laser fluorescence. Our observations, which were quali-tative in nature, indicate that this line has a width on the order of 0. 002 nmFWHM at very low NO2 pressure. However at pressures of a few mTorr thefluroescence intensity displayed little variations as the laser was tuned overthe width (0. 01 nm) of the gain curve. Thus we are led to suspect that NO2

has an anomalously large self-broadening cross section, and in consequence,observation of its absorption line structure in the visible requires very lowpressure, and long path length measurements. The contrasting ability toisolate individual strong lines in the I2 spectrum is an advantage for studiesof tuned laser fluorescence and the approach to resonance. Our work withI2 has included the following experiments, of which the last two have beensupported on this contract:

Measurement of absorption within the 514. 5 nm gain curve. - Thesemeasurements involved the argon laser, single-moded and tuned by tiltedetalon. The results, shown in Fig. 15, were obtained to evaluate the useof an I2 filter to block the exciting line in Brillouin spectres copy.

Measurements of re-emission intensity as a function of separation fromresonance and background gas pressure. - The purpose of these experimentsis to determine whether re-emission from I2 excited slightly (~0. 005 nm) offresonance is characteristic of scattering or fluorescence. The determina-tion involved repetition and extension of the experiments of Fouche and Chang(ref. 13). Our conclusion is that the re-emission shows nearly counter-balancing effects of line broadening and quenching. The theoretical analysisof these effects in Section II accounts satisfactorily for the relatively weakdependence of re-emission intensity on background gas pressure observed

56

Page 62: study of resonance light scattering for remote optical probing

1.0

0-9

0.8

0 0.7

§ 0.6

1 °'5cr*- 0.4

| 0.3

o 0.2

8: o.i

40°Ci.o0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

o.i

I I I I I I I I I-0.06 -0.04 -0.02 0 0.02 0.04

WAVELENGTH SHIFT IN ANGSTROMS FROM CENTEROF ABSORPTION LINE

Figure 15 Absorption of I2 vapor as a function of laser wavelength withinthe 514. 5 nm gain curve of an Ar laser. The laser is single-moded andtuned by a tilted etalon, with wavelength monitored by a Fabry-Perot inter-ferometer. Laser power into cell was 10 mW and laser linewidth < 0. 0001nm (0. 001A). The I2 vapor pressures correspond to the indicated cold-finger temperatures.

508 509

WAVELENGTH. NH

510

Figure 16 Antistokes fluorescence from iodine excited on a line within the514. 5 nm Ar laser gain curve. The two main peaks emanate from theoriginally excited I2 state, while the satellite (collision) peaks result fromenergy exchanges in collisions associated with ±2 changes in rotationalangular momentum quantum number.

57

Page 63: study of resonance light scattering for remote optical probing

by Fouche and Chang, and by St. Peters et al., (ref. 14). Because collisioneffects are still significant and particularly evident in the spectral finestructure of the re-emission at the largest separations from resonance in-volved in these experiments, we believe that the re-emission is moreaccurately described as fluorescence than as scattering.

Measurement of collision to main peak intensities as a function ofseparation from resonance. - This experiment was undertaken to determineif collision rates depend on separation from resonance. The antistokesfluorescence spectrum of 1% excited within the 514. 5 gain curve is shown inFig. 16. The main peaks form the strong doublet resulting from direct re-emission from the excited state, whereas the smaller peaks result fromcollisions imparting a change in angular momentum in units of A J = ±2 beforere-emission. Theoretical results for the time dependence of re-emissionled us to speculate that as the incident light is tuned away from resonance withthe transition producing this spectrum, at some point the change fromfluorescence to scattering should begin, evidenced in this case by a reductionin the collision to main peak intensity ratio. The results obtained upon tuningbeyond several Doppler widths is shown in Fig. 17, for the case of oneresonance line within the Kr ion laser gain curve at 568. 2 nm. It is evidentthat the effect described above was not observed. In the case of anotherresonance line near one edge of the 514. 5 nm gain curve, we were able toobserve the collision to main peak ratio out to more than 1. 4 GHz from theresonance center on one side, which is about 3 times the Doppler width.While the two peak intensities decreased by a factor of nearly 104, theirratio again remained constant to within a small experimental error.

There are several plausible explanations for the failure to observe the on-set of a change from fluorescence to scattering. Among these are:

1. The speculations drawn from the classical theory are wrong. Inparticular, there may be quantum mechanical effects which areimportant even to the qualitative nature of the phenomenon as observedin these experiments.

2. These may be significant inhomogeneous broadening, other thanDoppler broadening, which extends to the wings of the absorption line.In this case, the molecules responsible for most of the re-emissioncan be those momentarily shifted into resonance by the fields of nearneighbors. *

With respect to the first point, two of our group (S. D. Silverstein andR. L. St. Peters) have developed a quantum mechanical calculation whichsuggests that a constant collision-to-main peak ratio may not be unexpected

58

Page 64: study of resonance light scattering for remote optical probing

-£ 100-CO

COOr<t

oo:LLJl—LU

Oct:

o.CO

CO

COCO

LU

I I I I I I I I

o MAIN PEAK

I I I I I I I I I I I

• NORMALIZED FIRSTCOLLISION PEAK

TUNING STEPS 1.2 x I0~4nm

Figure 17 Normalized collision and main peak intensities as a function ofseparation from resonance.

59

Page 65: study of resonance light scattering for remote optical probing

even in the absence of inhomogeneous broadening. This calculation, however,involves some subtle points which have not yet been adequately resolved.

Whatever the case, the persistence of the collision peaks is strong evi-dence that this re-emission is fluorescence rather than scattering, as thoseterms are defined by Holzer, Murphy and Bernstein (ref. 4), and as definedwithin this report.

Measurements of depolarization of re-emission excited above and belowthe dissociation limit. - These experiments were conducted to verify thetheoretical: calculations "of depolarization discussed inSection II. The experi-mental conditions and results have already been described in that section.

Measurement of the Ozone Cross Sectionat Visible Light Wavelengths

The cross section for O3 vibrational Raman scattering (vt - 1110 cm"1)was measured for this contract in order to establish a benchmark for cor-responding UV measurements. The ozone was generated in a flowing streamof O2 by photo-dissociation; using a low pressure Hg discharge lamp speciallydesigned to maximize output at the 184. 7 nm line relative to the 253. 7 nmline. The lamp was enclosed in a 15 cm I. D. pyrex pipe with aluminum endplates and Teflon polymer gaskets. Oxygen pressure and flow rate weremeasured with suitable precision instruments. The O2/O3 mixture wastransported through 3/8" I.D. Tygon tubing to the first of two Spectrocilquartz cells (Markson cells, 5 cm long, 2 cm I.D. , with two ports) whereO3 concentration was measured by transmission of 253. 7 nm radiation froma low pressure Hg lamp. Parallel interference and monochromator filterswere used to ensure that only 253. 7 nm was observed. The slow decay of O3

in the cell (~25$ in 10 minutes) under non-flowing conditions indicated thatthe 253. 7 nm radiation level used for concentration measurements was notsufficient to significantly alter the O3 concentration at the flow rates employed.(The 253. 7 absorption in O3 is dissociative.) Ozone concentrations werecalculated using well established absorption coefficients (ref. 19). Thesemeasurements indicated that the O3 generator is a stable source providing O3

concentrations which can be controlled easily by varying the flow rate . Theconcentration obtained at an O2 flow rate of 400 cm3/min is 3. 2 torr O3 in 740torr O2.

The second quartz cell was connected to the first cell by 75 cm of 0. 95cm (3/8") Tygon tubing. This cell was used for the scattering measurements.The cells were switched with tubing connections unchanged to determine ifthe O3 concentration was significantly changed while flowing through the Tygontubing. After flow at 400 cm3/min and O3 generation had been established forone hour, the concentration in the second cell stabilized at 95$ of that in thefirst cell.

60

Page 66: study of resonance light scattering for remote optical probing

The gas mixture exiting the second cell was passed through Linde Type13X Molecular Sieve to convert O3 back to O2, and then exhausted into avacuum system through a metering valve.

The scattering intensity of O3 was measured relative to that of C^ in thestandard 90° geometry illustrated in Fig. 10. Depolarization of the O3 scat-tering was found to be less than 20$. The techniques used in these measure-ments have been described in detail in a previous publication (ref. 40). Witha 0. 5 nm FWHM semirectangular slit function, we found that the ratio of O3 to Ojvibrational Raman scattering, at jshifts of 1110 and 1557 crn~ l, respectively,is 1. 96 for 514. 5 nm excitation and 2. 05 for 488. 0 nm excitation. The esti-mated error in these measurements is ±20$. The predominant sources ofuncertainty are the concentration and scattering count rate determinations.Using our previously measured value for the O2 Raman scattering crosssection, we calculate for the absolute O3 vibrational cross section a value of1. 01 x 10~30 cm2/sr ± 25$, at 514. 5 nm, for the scattering geometry of Fig. 10.

Fouche and Chang (ref. 41) have reported the O3/N2 vibration crosssection ratio to be 4. 0. In another publication (ref. 42) they report the cor-responding O2/N2 ratio to be 1. 2, which is in agreement with our result.Thus their results lead to the moderately larger value of 3. 3 for the O3/O2

vibrational RS cross section ratio at 514. 5 nm.

Spectrometer-Dye Laser Facility

The Spectrometer-Dye Laser Facility (SDLF) includes a Spex 3/4 meterspectrometer/monochromator and a pulsed dye laser whose output can bedoubled, producing incident light over the wavelength range from 230-760 nm.The dye laser, which is pumped by an AVCO Model C950 pulsed N2 laser,provides pulses with spectral width of 0. 005 nm and time width of 5 nsec. , at arate of 100 pps. The average power in the visible is lOmW, and in the uv(doubled), 10-100 uW. Photomultipliers and pulse-counting and timing dataanalysis are incorporated into this facility, which provides time resolutionof 6 nsec. Concurrently, the monochromator provides spectral analysis ofre-emitted (or incident) light to 0. 02 nm. A Fabry-Perot etalon is used tomonitor the dye laser output to higher resolution. The system configurationis illustrated in Fig. 18. A detailed description of components of the systemfollows.

Dye laser. - The dye laser is pumped by a pulsed N2 laser; its configu-r at ion is similar to that described by Hansch (ref. 43). The pulsed N2 pumplaser emits pulses of 100 KW peak power at 337. 1 nm, and has an averagepower of 100 mW when operated at its maximum repitition rate of 100 pps.This beam, which is rectangular in cross section with < 2 mr divergence inthe short dimension and < 30 mr divergence in the long dimension, is broughtto a near focus just inside the surface of a flowing dye cell with an optical

61

Page 67: study of resonance light scattering for remote optical probing

ECHELLE GRATING ANDROTATIONAL STAGE

I IFROM AVCO C950 N9 LASERi i '

' !LENS /PARALLEL PLATE POLARIZERREFLECTOR

RT. ANGLE REFL.

D""'STEPPINGMOTOR

EXPANDER \ APERTURE

FLOWINGDYE CELL

STEPPINGMOTOR DRIVEAND READOUT

WAVETEKGENERATOR

SHG XTAL

FILTER c±=>/I /i/

8%'R

DIRECTIONSWITCH

LENS

DIFFUSE REFL.

PHOTODIODE

-TO H.V.ELECTROMETER

SCATTERINGCELL

RT ANGLE REFL.

LENS

POLARIZATION SCRAMBLER

MONOCHROMATOR

TO CHARTRECORDER

• TOPM.DETECTOR

Figure 18 Tunable dye laser and optics for scatteringexperiments.

62

Page 68: study of resonance light scattering for remote optical probing

grade fused quartz lens of 5" focal length (FL) and 3" diameter. The focalregion is a narrow line about 1 cm long and about 0. 3 mm high. Most ofthe N2 laser beam energy is absorbed within a fraction of a mm just insidethe dye cell. The concentration of the dye is set high (about 5 x 10 ~3 Mfor rhodamine 6G) to produce strong absorption and a narrow rectangularactive volume.

A detail of the flowing dye cell is shown in Fig. 19. The cell was con-structed by fusing two fused quartz tubes together as shown in the side viewcross section. Quarter inch fused quartz tubing was then attached to theouter cylinder at the angles shown to produce a smooth flow at the input andoutput connections. The cell was then cut at an angle of 8° to 12° as shownin the top view of Fig. 19. Good optical quality windows purchased fromLambda Optics, antireflection (AR) coated with a broadband multilayer coat-ing on one surface by Broomer Research Corp. , were then carefully epoxiedon the sides of the 1" diameter tube with the coated surfaces to the outside.Care was taken to avoid blocking the optical path just inside the quartz tubewall with the cementing epoxy. Epoxy was also placed on the inner tube toseal the chamber from the dye solution. The 8° to 12° angle prevents re-flection from the windows from being amplified along with the laser beam tosignificant proportions by the extremely high gain active volume. Apump with a maximum flow rate of 6. 3 gal. per minute at 5 psi, speedcontrolled by a variable voltage transformer, is used to circulate the dye sol-ution through the cell. The dye cell was rigidly mounted on a Brinkman modelMP-ll/R micromanipulator that allows 3-D translational positioning of thecell with respect to the dye laser optics while the system is operating.

The tuning component of the dye laser is a 58 x 58 mm, 300 ^echelle grating blazed at 61°. This grating is mounted on an Aerotech ModelATS-301R rotational stage. The Aerotech stage has positioning resolutionof 0.1 arc sec when using the fine adjust differential screw. An AerotechATS- 30 ID digital stepping motor and readout is used to rotate the grating.Since the differential screw on the rotational mount moves in and out, thestepping motor used to drive the screw had to be mounted on a translationalstage that was free to move back and forth with the linear motion of the dif-ferential screw. A ceramic insulated coupler was used to couple the steppingmotor shaft with the differential screw shaft. The ceramic coupler providesthermal isolation between the stepping motor and the echelle grating. Thegrating can be slewed at a fixed rate or stepped manually with the Aerotechdrive unit or can be stepped at a present constant rate from an externalWavetek model 115 VCG generator. The direction of rotation can also becontrolled externally. A digital readout is provided by the Aerotech driveunit and serves as a reference for positioning the grating.

63

Page 69: study of resonance light scattering for remote optical probing

GROUND SURFACE FOR ATTACHMENTOF WINDOWS WITH EPOXY

111/16"

FIRE POLISH ENDS

1/4" CFQ

Figure 19 Flowing dye cell for N2 Laser-Pumped Dye Laser.

To fully utilize the 58x58 mm grating a beam expander is used to expandthe cross section of the light beam emitted from the thin lineal active regionby a multiple of 51 times. This is accomplished with a 3. 9 mm FL by 1. 1mm dia. input lens and a 200 mm FL by 50 mm dia. output lens. The lensesare near diffraction limited and AR coated for high transmissivity in thevisible spectrum.

On the opposite side of the optically excited dye region is a parallelplate polarizer that consists of 10 cover glass slides placed at Brewster'sangle and mounted in a rotatable mount that allows the linear polarizationof the laser to be set at any desired angle about the laser beam. One surfaceof a 2" dia. glass plate that is slightly wedged and mounted in a Lansing AODmount is used as the output reflector. The single surface reflector gives a 4$reflection and is quite adequate for the high gain achieved by the rhodamine6G dye. The echelle grating and rotational mount, the beam expander, theparallel plate polarizer, and the output reflector are mounted on a 6" x 36"invar surface plate for good temperature stability.

64

Page 70: study of resonance light scattering for remote optical probing

The Ng laser-pumped-dye laser system operates in the manner of ahigh gain two pass amplifier, rather than as a typical laser which builds upcavity radiation modes by amplifing multiple reflections between resonatingreflectors (ref. 43). The oscillation is limited to the two passes made by theshort (10 nsec) pump pulse, and the round trip transit time of the lasercavity. The small lineal active volume, which has a gain of about 103,amplifies the spontaneous emission, producing an observable amount ofsuperradiance along its axis. The 4$ of the superradiance which is reflectedby the output mirror is reamplified 103 times when passing back through theactive region. This signal is then expanded 51 times to fill the echellegrating. The retroreflected light from the grating is angularly dispersed andonly a small fraction of the original light, that lying in a narrow angular (andtherefore narrow spectral) range is recollimated by the beam expander andsent through the active region for its second amplification and passage throughthe output reflector. A 4 mm diameter aperture, placed after the output re-flector, passes the laser beam and blocks'most of the unwanted light whichoriginates from the initial superradiance and from reflections off windowand lens surfaces.

The spectral width of this system is determined entirely by the angulardispersion of the grating and beam expander combination and is limited, ofcourse, by the maximum resolution of the exposed grating surface which inour case would be about 4xlO~3 nm at 600 nm.

Laser performance characteristics. - The maximum average powerobtained with the N2 laser-pumped-dye laser using a 5 x 10~3 concentrationof rhodamine 6G, was a little better than 10 mw as measured with a CoherentRadiation Model 205 thermopile. For this power level a Bausch and Lomb1200 gr/mm grating was used in first order in place of the 300 gr/mm echellegrating and there was no polarizer in the laser. In this instance the emissionlinewidth integrated over several hundred laser pulses was about 4 x 10~2 nm,as estimated from Fabry-Perot interferograms. When the echelle gratingwas used (10th order) and the parallel plate polarizer placed in the lasercavity, .the average power dropped to the 3 mw range. However, with thisgrating, the laser linewidth is about 5 x 10~3 nm, which is near the theoreticallimit for the number of grooves exposed.

If the 4$ output reflector is replaced with an uncoated fused quartz etalonof 14. 7 mm thickness (FSR - 8. 4 x 10"3 nm), the laser emission width narrowsfurther to about 2 to 3 x 10~3 nm. With the etalon output reflector, however,the spectral output appears to "mode hop" as the laser is tuned by rotatingthe grating. Sometimes two laser emission lines appear at a separationcorresponding to the free, spectral range (FSR) of the etalon resonant reflec-tor (8 x 10~3 nm). In some instances the "mode hopping" of the laser pro-duces a 30$ periodic variation of the laser output when the laser wavelengthis continuously scanned. With the 4$ output reflector, though, a single laserwavelength could be continuously scanned with no alteration of the output power.

65

Page 71: study of resonance light scattering for remote optical probing

The N2 laser-pumped-dye laser has also been operated with the dye 4methylumbelliferone (4-mu) in perchloric acid and ethanol solution (ref. 44).This dye forms a strong lasing exciplex (4 mu) H* in the acidic solution andlases from 460 nm to 560 run depending on the perchloric acid concentration.The optimum concentration for laser output in ethanol is from 10~2 to 1. 5 x10~2 M of 4 mu and 1 to 3 M of perchloric acid with the 1 M perchloric acidconcentration optimum for the green wavelengths and the 3M acid concentra-tion for the blue-violet wavelengths. Typical average output power of 5 mwwere achieved with the 1200 gr/mm grating in first order.

The angular divergence of the dye laser beam with the R6G dye wasestimated to be 1. 2 mrad (half angle) for the central bright spot. Thisdivergence allows one to estimate the diameter of the active volume as 0. 25mm for a diffraction-limited wave. The 4 mu dye laser beam cross section,however, was an elongated ellipse and estimations of the half angle divergencegave 1 mrad for the vertical direction and about 12 mrad for the horizontaldirection. The larger divergence in the horizontal direction indicates thatthe N2 laser pumping radiation is absorbed within a depth of about 30 urn inthe dye cell.

Second harmonic conversion. - A right angle dielectric reflector, coatedfor maximum reflectivity in the yellow-orange part of the spectrum, directsthe dye laser beam through a 10 cm FL lens that brings the laser beam to afocus inside a 20 mm cube, ADP crystal whose z axis is oriented at 62° forfrequency doubling at 590 nm. The ADP crystal is mounted in a Lansing AODmount and by angulating the crystal relative to the laser beam the doublingfrequency can be angle tuned a few tens of nm about the center wave-length of 590 nm. A calibrated Corning CS 7-54 filter is used immediatelyafter the ADP crystal to completely block the visible laser beam and passabout 85$ of the second harmonic.

A 5$ conversion efficiency in frequency doubling to the uv is easily ac-complished with the ADP crystal. After the losses of the filter and re-flections losses from lenses and a right angle reflector are sustained, weeasily achieve 100 nW average power at the scattering cell with 5 x 10~3 nmlinewidths. The uv laser power is monitored using a slightly wedged quartzdisk as a single surface reflector to direct 4$ of the beam toward a diffuse(BaSO4) scattering surface (cf. Fig. 18). A known solid angle of the scat-tered laser light is then detected by a calibrated ITT S-5 vacuum photodiode.The pulse energy is found to vary by 20 to 30$ from laser shot to shot.However, the average power varies by no more than about ±3$ from a meanvalue over a second or two of averaging. The large shot to shot variationin the laser pulse energy is probably due to thermo-optic distortions

66

Page 72: study of resonance light scattering for remote optical probing

created by hydrodynamic instabilities in the fluid flow through the activeregion of the dye cell. These effects average out, however, over severalhundred laser shots.

Two other crystals were used for second harmonic generation of thevisible dye laser light. A KDP (potassium dyhydrogen phosphate) crystalwas purchased from Quantum Technology, Ltd. This crystal was 45° z cut for90° (noncritical) phase matching near 595 nm at room temperature. The 90°phase matching angle can be temperature tuned to about 602 nm by heating thecrystal to 100°C. With noncritical phase matching the SHG efficiencies shouldbe greater than for critical phase matching (i. e., angle tuning) since there"isno Poynting vector walk-off for the noncritical matching case. Using a 5. 5cm F. L. lens to focus the 10 mW dye laser beam into the KDP crystal, uvlight was generated when the laser wavelength was set on the proper phasematching wavelength for room temperature. The KDP crystal temporarilygenerated about 100 nW of UV power but quickly damaged with the generationof an inclusion in the crystal where the laser beam was focussed. In thisinstance the laser power density was clearly above damage threshold of thecrystal. Estimates of the laser peak power density with the 5. 5 cm F. L.lens give over 500 Mw/cm2, a value that can damage most crystals.

The third crystal used for SHG was lithium formate monohydrate (LFM).This crystal was first used by Singh et al. (ref. 45) for frequency doublingthe 1. 06n Nd lasers. Lithium format is of interest to us because it can beangle matched down to its absorption band edge near 230 nm. Our 1 cm3

LFM crystal was purchased from Isomet and was protected from dehydrationby immersion in a Dow-Corning silicone fluid and placed in a sealed containerwith an AR coated window for the visible input laser beam and a fused quartzoutput window for the SHG beam. The crystal was cut at 43° to the z-axisto allow angle phase matching at 275 nm with the laser beam at normalincidence to the crystal faces. Table II shows the results of our SHG ^measurements with LFM from second harmonic wavelengths of 250 nmto 230 nm. The measured phase matching angles (8p meas.) agree quite wellwith our computed values (0p comp) that were made from a projection ofSingh's data in the ir and visible spectrum. The LFM crystal, however, wasdamaged by the optical radiation at all the above listed second harmonicwavelengths. It was found, for example, that at 250 nm the output powerof the SHG radiation fell 5 to 10$ after 10 min. of operation and a small whitecloudy scattering center developed at the position of the laser beam focus inthe crystal. This situation became more severe as we proceeded to the shorterwavelengths. The absorption of the uv radiation by the crystal becomes largerand this in turn causes a local temperature rise at the focus of the laser beamin the crystal that can become large enough to dehydrate the crystal and pro-duce the white cloudy inclusions observed in the crystal. If the damage isthermo-optic as just described then the crystal or laser beam could be rotatedto avoid overheating and consequent damage.

67

Page 73: study of resonance light scattering for remote optical probing

TABLE II

, . n SHG Conversion Efficiency With(nm) 6 meas. Q comp. , . „_ . _ -_,•; ic 10 cm FL lens 8 cm FL lens

250 40.6° 39° 3.1$ 2.0$

245 39.8° 38.5° 2.3$ 1.9$

240 38.9° 37.5° 2.3$ 1.3$

235 37.9°. 36.7° _...' 1.7$_ .. 0.9.$

230 36.3° 35.6° . --- 0.3$

Spectrometer and Related Optics

The horizontally polarized UV beam generated by the ADP crystal passesthrough a 150 nm FL lens, and is then redirected vertically upward by aright angle, prism. The beam, is brought to a focus at the center of a 5 cm dia.by 2 cm wide cylindrical scattering cell (cf. Fig. 18). The focus of the UVbeam is centered 20 cm away from the entrance slit of the Spex model 1800,'3/4 meter, f 7. 5 monochromator. A 50 mm FL quartz lens, held in trans-latable mounts and placed between the scattering cell and the monochromatorslits, collects the light scattered at right angles from the laser beam,making a 1:1 image of the laser beam on the entrance slit of the monochro-mator. A crystal quartz polarization scrambler is placed ahead of theentrance slit to eliminate any polarization sensitivity of the monochromatorto the incident light to be analyzed. The Spex monochromator has a 1200gr/mm grating blazed at 300.0 nm and is used in first order. The resolution ofthe monochromator in first order is about 0. 02 nm.

The wavelength of the monochromator was calibrated at seven differentpoints in the uv using a low pressure Hg arc placed so as to illuminate theentire entrance slit and grating of the monochromator. The relative spectralresponse of the monochromator from 250 to 400 nm was also measured usinga calibrated 200 watt GE quartz iodine lamp and a calibrated 6 1/2" squareCorning CS 7-54 filter. The purpose of the filter is to reduce visible lightfrom the source, which otherwise scatters within the spectrometer, causinga high background. Remnant background from visible light, plus the darkcount, was evaluated using a Corning CS 0-54 filter which transmits lightonly at wavelengths longer than those of interest. The standard lampilluminates a diffuse scattering surface constructed of Eastman high reflect-ance paint which is placed in front of the monochromator along the optic

68

Page 74: study of resonance light scattering for remote optical probing

axis of the collecting lens. The single photon photomultiplier pulses werecounted in a specified time interval at many different wavelengths for thecalibration. Signal linearity was checked by using uv calibrated neutraldensity filters.

Detection and Time-Analyzing Electronics

Figure 20 shows a schematic of the detection system used for either aspectral analysis or a time analysis of the scattered light. The output of themonochromator is sent to a EMI 9635 QAM Photomultiplier (PM) speciallyselected for low dark noise. The -PM is mounted in aProducts for ResearchTE 200 housing that is mounted at the exit slit of the monochromator. Thelight levels required for this system must be low enough so that we obtaineither one or no detected photons for each laser pulse. Each detected photonpulse is amplified by a Kiethley Model 109, 50 Q input, 20 db pulse amplifierthat is placed at the output of the PM housing. The amplified pulses then goto an Ortec Model 463 constant fraction discriminator (CFD). The singlephoton pulses from the amplifier were monitored at the input of the CFD withan oscilloscope and found to vary in pulse height from about -50mV to -SOOmVwith pulse widths of about 25 ns. The electrical noise pick up from the N2laser, on the other hand, was only a ±6mV signal. The CFD, which has aminimum input level of -50mV and therefore will not trigger on the electricalnoise, is triggered to give an output pulse at a constant fraction of the inputpulse height rather than at a constant input level. Triggering on a constantfraction of the pulse height substantially reduces the timing error (walk) thatis caused by the varying single photon pulse heights generated by the PM.The output pulse from the CFD is sent to the stop input of an Ortec Model437A time to pulse height converter (THC).

Initially, a Berkley 903 double pulse generator sends out two triggeringpulses whose relative separation in time is adjustable. One pulse initiatesthe N2 laser and the other pulse is sent to the start input of the THC. At alater time, determined by the time interval set on the double pulse generator,the delay in the triggering electronics, and the time response of the scatter-ing species, the detected photon signal from the CFD may enter the stopinput of the THC. If this pulse is within the time window viewed by the THC,it generates a pulse output from 0 to 10V whose height is proportional to thetime interval between the start and stop input signals. The output of the THCis coupled to a Northern Scientific NS-710 pulse height analyzer (PHA). Afterseveral thousand laser pulses a time response of the scattering species isbuilt up in the appropriate channels (or bins) of the PHA.

By counting a small range of pulse heights from the THC the detectionsystem effectively time gates and consequently discriminates against unwantedbackground pulses that occur randomly in time. The minimum full scale time

69

Page 75: study of resonance light scattering for remote optical probing

COUJ co•1-1

•4-iO

co a:= 00,5

fy*|« i • . ti-i-j LLJ

•~™ 3SCD UJ

0°Q

CDfV*

I—QCUJ 1t/°>J2—i

5 E^»— =d<*co p^-

fc*- - O-^ 11

I—nGL^Tpe>

CD fe

oc _ i pUJ Ij U- .

r *— LU °r ^o0— Of<» ,s

^?*—-»• < UJoc: i— jg

* UJ CDUJ

«=4^b era: <=» &^>^ u. o o ^co1--,- fe Si^^ ' ^^ *lj

^ * S Uli 0 1-i a: uj

31 QO0 2z a.

O 2g cOuj •*= w

Q_^ oc:2^£ 3 -CO UJ^ >--1 Q

CM

0)-M0

•a<_n

•rH

^,£CQ

^^fa_]

>»-l-i•H• r-l

On!

0COrtJ

Qd a;

feS «**; iv* ^

zo |og g

^ 5oCO

OJX!-4-1

<t-l

O

S ra

Fig

ure

20.

Blo

ck d

iagra

and a

naly

sis

ele

ctr

onic

70

Page 76: study of resonance light scattering for remote optical probing

range over which the THC can be set to detect a PM signal is 50 ns. Thisrange is divided up into 1024 channels in the PHA or 50 ps per channel.

The output of the THC is also sent to a Tennelec TC 592P digital rate-meter which counts all the pulses from the THC regardless of height providedthe lowest pulse height is set above the ratemeter threshold of 0.15V. Thissituation can be easily obtained by adjusting the time separation of the twopulses from the double pulse generator to give THC output pulses that lieabove the ratemeter threshold. In this mode the THC serves as a coincidencegate. The digital ratemeter has an analog-output that drives one pen of adual pen strip chart recorder. The other pen on the recorder is driven bythe output of the Kiethley electrometer that monitors the laser power. Byscanning the laser wavelength at a fixed monochromator wavelength or viceversa, the scattering signal can be recorded as a function of laser wavelength(or monochromator wavelength) along with the average laser output power.

The THC and PHA were calibrated by using the two pulses from thedouble pulse generator to start and stop the THC. A dual beam oscilloscopemonitored the start and stop input pulses. By setting the time interval be-tween the pulses from the double pulse generator to different positions we

.Calibrated the PHA with respect to the oscilloscope time scale.

By observing the number of PHA channels required for counting at fixedtime interval of the double pulse generator, we could determine the jitter inthe interval between the pulses to be 1 ns. This represents one limitation onthe time resolution capabilities of our detection system. Figure 21 shows theresults of data taken of the Rayleigh and Mie scattering from the laser beamin air. Each data point in Fig. 21 corresponds to a channel of the PHA. Inthis instance the THC range was set on 1 p.s. Since there are 1024 channelscovered by the 1 p.s range the separation between each channel correspondsto about 1 nsec. Therefore, the half width of the pulse in Fig. 21 is about6 ns. This width represents the average of many laser pulse widths plus thejitter in the triggering electronics as discussed above and any jitter in theinitiation of the laser action. Thus, the time resolution of the system isdetermined to be 6 nsec.

Signal Counting Statistics

The maximum pulse repitition rate of the dye laser is limited by that ofthe pump laser to 100 Hz. In our experiments we usually ran the laser atthe slightly lower frequency of 94 Hz, which is set by the double pulsegenerator. In this case the maximum signal rate that can be indicated by thedetection system is 94 Hz. However, the indicated count rate begins todeviate significantly from true count rate below 20 cps. This behaviorreflects the probability that more than one photon will be detected during theTHC "on" period, since the THC responds only to the first detected photon.

71

Page 77: study of resonance light scattering for remote optical probing

nSEC

Figure 21 Time dependences of Rayleigh/Mie scatteringas displayed by the multichannel "analyzer 7" This re-suit indicates the time response of the system to an"instantaneous" process.

The relationship between indicated count rate CT and true count ratecan be determined from a simple statistical analysis, which yields

= Rln

where R is the laser pulse rate. The ratio C^/Cj is plotted in Fig. 22.Experimental data presented subsequently are corrected by this factorwhenever appropriate.

Absolute N2 Vibrational Raman ScatteringCross Section For Incident Light at 300. 0 nm

This cross section for the vibrational RS from N2 was measured usingthe SDLF in order to provide a standard against which to measure other crosssections (and effective cross sections). A typical RS signal from N2 is shownin Fig. 23. The cross section value was obtained by measuring the ratio ofRaman scattering to Rayleigh scattering at several different gas pressures(nominally 300, 500 and 700 Torr) and then multiplying this ratio, correctedfor spectrometer response, by the theoretical value of the Rayleigh crosssection, given by

THRay '

Here n is the refractive index measured at a molecule number density of N.A small depolarization correction (ref. 18) is omitted in this equation. Usingn-1 = 3. 13 x 10"4 for N2 at 300 nm and STP, obtained from the InternationalCritical Tables (ref. 46), we calculate

o™ (300 nm) = 6.60 X 10"27 cm2 sr'1

72

Page 78: study of resonance light scattering for remote optical probing

1.6

1.5

oO 1.4

COCQO

1.3

<ta:

1.2

.00 O.I 0.2 0.3 0.4 0.5 0.6

(OBSERVED COUNT RATE) / ( L A S E R PULSE RATE)

Figure 22 Relationship between true count rate and indicatedcount rate for the MDLF.

73

Page 79: study of resonance light scattering for remote optical probing

M

~ 4

oo

oUJ

oUJ»-UJ

N2 RAMAN SIGNAL

322.0 322.5 323.0

MONOCHROMATOR WAVELENGTH ( nm )

Figure 23 Typical signal obtained for vibrational Raman scat-tering from N2 at 740 Torr.

74

Page 80: study of resonance light scattering for remote optical probing

The ratio GRAM/aRAY was measured on two separate occasions in the SDLFusing wide slits and a technique very similar to that described in ref. 40. Theresulting values for this ratio were 1.50 x 10~3 and 1.44 x 10~3. An averageof these results yields

aRAM(N2) = 0.97 x 10-29 cm2/sr ±25$

This cross section is for the Q-branch vibrational RS from N2 summed overscattered light polarization in the right angle scattering geometry of Fig. 10.It is interesting to note that a (1/X2)4 extrapolation of the value for thecorresponding cross section measured at 514.5 nm (ref. 5) yields0. 463x10 29cm2. Thus the measured value is about 2.1 times larger than theextrapolated value. We believe that this weak enhancement is real. Indeed,the form of quantum mechanical expression for the cross section, Eq. (38),suggests that the cross section should increase faster than (1/X2)4 over thiswavelength interval because of the increased proximity to the strong vacuumUV resonances of N2, which begin near 140 nm. The Rayleigh cross sectionalso increases faster than(l/Xa)4- because of the dispersion of the refractiveindex, but the increase in the Raman cross section is more pronounced.

The O2 Vibrational Raman Cross Sectionfor Incident Light at 300 nm

We also measured the ratio of O2 to N2 Q-branch vibrational crosssections for incident light at 300 nm. This result yielded 1. 59 ±20$ for theratio, up from 1.24 at 514.5 nm for the cross sections summed over polariza-tion. Thus the O2 cross section increases faster than the N2 cross section,which seems reasonable since the strong O2 uv resonances begin at a longerwavelength (~230 nm).

Laser-Excited Fluorescence from SO2

Sulphur dioxide exhibits relatively strong absorption between 260 and 320nm, and much weaker absorption at longer wavelengths around 380 nm. Aspart of the work supported by this contract, we have examined the re-emissionfrom SO2 following narrow band (.005 nm), pulsed (Snsec) excitation atvarious wavelengths near 300 nm. We have observed the low pressurefluorescence and phosphorescence spectra, and the sensitivity of features ofthis spectra to exciting wavelength, self quenching and quenching by N2 andair. In particular, we have found excitation wavelengths when the fluorescenceshifted by the vx vibration frequency is more than 104 times stronger in airnear STP than vibrational Raman scattering from N2. This fluorescenceappears as a sharp peak rising over background continuum radiation. Thesensitivity of the integrated peak intensity is such that it varies irregularlyover a factor of two as the laser is tuned over a range of about 0.1 nm.These results are described in more detail below.

75

Page 81: study of resonance light scattering for remote optical probing

Low Pressure SO2 Fluorescence

The absorption spectra of SC>2 has been studied and analyzed to someextent by several investigators (refs. 20, 47-53). To our knowledge N.Metropolis (ref. 51) has given the most accurate vibrational analysis of theground to first excited singlet state transition bands in the 260-320 nmregion. There is still some disagreement (ref. 53), however, over his assign-ment of the vibrationless (0, 0, 0)"-(0, 0, 0)' transition at 337. 6 nm. A. J.Merer (ref. 52) has given a rotational analysis of the much weaker band sys-tem at longer wavelengths around 380 nm and demonstrated that these bandsderive from a singlet-triplet transition.

The absorption spectrum of SO2 near 300 nm has been shown in Fig. 9.The labeling of the bands A, B, C,. . . is due to Clements (ref. 47) who assignedthese regularly spaced bands to a simple progression in the bending modevibrational quantum v2. Metropolis, however, demonstrated that these bandsare really a superposition of various higher combination bands whichfortuitously create the appearance of a simple progression. The rotationalsubstructure of these combination bands has not been analyzed to our knowl-edge. It is considerably complicated by the increased asymmetry of themolecule when excited to the first singlet state. Metropolis estimates a bondangle change from 120° to 100° when the molecule makes a transition to thefirst excited singlet state. The angle change is also accompanied by a changein the bond distances which eliminates the two-fold rotational symmetry thatthe molecule has in the ground electronic state.

For the SO2 fluorescence measurements the scattering cell was filledwith anhydrous grade (99.98$) SO2 at pressures of a few tens of mTorr. Wetuned the laser, whose bandwidth was 0. 005 nm, through portions of the J,H, G, and F bands shown in Fig. 9 and examined the fluorescence output atthe different laser wavelengths by scanning the monochromator from the laserwavelength out to longer wavelengths. In general the spectrum looked some-thing like that reported by Mettee (ref. 20) who used a much broader Hgemission lamp for an excitation source. In our case a large continuumfluorescence occurs which reaches a maximum at about 340 nm. Super-imposed on this continuum are many peaks whose positions relative to theexciting wavelength correspond to fundamental, overtone and combinationtone vibrational levels in the ground electronic state. At certain laser wave-lengths, however, it was observed that the fluorescence peaks became muchstronger than the continuum fluorescence. In addition, the fluorescencepeaks exhibited a shorter decay time than the continuum fluorescence. Aparticularly strong fluorescence peak whose wavelength was Stokes shiftedby the Vi symmetric mode vibrational frequency was observed at 299.98 nm±0. 04 nm, just to the short wavelength side of the peak of the G absorptionband. This strong fluorescence peak is shown in Fig. 24 along with several

76

Page 82: study of resonance light scattering for remote optical probing

orLUCO

CVJoCO

odoro

Oins

§X

O

O

10

ro

ac00O)

O5<M-<->rt

•*->•a

(1)

oCO

0

wCOa;

acoo

I2

o|

3

Ifa, ?a —.en .g4> QDt) CG v<U "—Io g|to 54) «>

O .

2 ^

0)^g)

0)4->O

0)IH

OCJ

77

Page 83: study of resonance light scattering for remote optical probing

other smaller peaks corresponding to overtone levels of the V2 bending mode,vibrational quantum and one peak that corresponds to a Vx + \^ combinationtone level. The background between the peaks is mainly due to a continuumfluorescence emission. The actual background with no SO% in the scatteringcell is considerably smaller than the continuum fluorescence. The spectrumfor Fig. 24 was taken with about 10 mTorr of SO% in the scattering cell and50 ^m monochromator slits. The overtone and combination tone structurecontinues further out to longer wavelengths not shown in Fig. 24 with anothermoderately strong peak at about 3vj (334. 7 nm). When the laser is tuned only0.08 nm away from the wavelength used for Fig. 24, the discrete spectrumfalls off drastically leaving only a reduced vr peak. The continuum . . - .fluorescence remains at about the same level, however. This is shown inFig. 25 where the laser wavelength was reduced to 299.90 nm.

The sensitivity of the discrete fluorescence to incident wavelength can befurther demonstrated by opening up the monochromator slits to obtain asemirectarigular passband of 0. 55 nm and monitoring the entire Vx fluorescencechannel while tuning the laser wavelength over a small range of about'0. 12nm. A tuned laser fluorescence spectrum of this sort from 1 Torr of SC>2is shown in Fig. 26; This data was obtained in the form of a histogramproduced by the digital ratemeter counting for 10 sec intervals while the laserwavelength was swept at a rate of 1. 16 x 10~4 nm/sec (i. e., one angular stepof the grating every 5 sec). To produce Fig. 26 the data was taken from thehistogram trace with a graphical digitizing system and placed on magnetic tape.Then the data was corrected for single photon detection by a computer, asdescribed in the previous section and traced out on a convenient size scaleby a digital plotter. The absolute wavelength measurements should be goodto within ±0. 04 nm. The laser wavelength timing per angular step of the lasergrating was measured by stepping the grating 1000 steps with the digital driveand readout devices and taking the difference between two wavelength mea-surements before and after the 1000 steps. From this measurement it wasdetermined that the tuning rate was 5.8 x 10~4 nm per step. Since theaccuracy of this measurement was only about 10$, the accuracy of the wave-length scale in Fig. 26 is only 10$ . The linearity of the scale is good toabout 0.1$ however.

From Fig. 26 we see that the fluorescence into the Vj channel shows agood deal of fine structure when the laser is tuned over this region of thespectrum. Strong .structure was also seen at other regions of the spectrum.For example, four strong peaks were observed when tuning the laser througha region of the J band near 296.08 nm and observing the fluorescence in theVi + 2v2 channel (217.6 cm"1) at 316.47 nm. A large part of the tuned laserfluorescence (TLF) spectrum, however, did not show a strong fine structure.It was felt that the TLF spectrum in Fig. 26 corresponds to absorptionbetween certain rotational levels in the two electronic states of the 803

78

Page 84: study of resonance light scattering for remote optical probing

79

Page 85: study of resonance light scattering for remote optical probing

CO

QQOC.

CD

CO

OCOLUCCO

100

_ 50

I299.85 299.90 299.94

LASER WAVELENGTH (nm)

300.00

Figure 26 Tuned Laser Fluorescence (TLF) spectrum of SO2. The mono-chromator slit function used to obtain this data is semi-rectangular withwidth 0. 55 nm, centered on the \>i fluorescence line.

80

Page 86: study of resonance light scattering for remote optical probing

molecule. To verify this speculation Dr. R. J. Exton at NASA LangleyResearch Center has taken the absorption spectrum of SO2 with a 2 m Czerny-Turner spectrometer in 3rd order with resolutions of 2. 3 x 10~3 nm and4.6 x 10~3 nm in the same region of the G band. The absorption cell usedwas 305 cm long (one pass) and filled with 50 mTorr of SOg. A high pressureHg lamp was used as the source light for the absorption measurements.Exton1 s absorption curves also show some structure in this region of thespectrum superimposed on a larger continuum absorption. In comparing thetwo sets of data, it was found that when the TLF curve was shifted by6.8 x 10~2 nm, which could allow for wavelength measurement inaccuracies,there was striking correspondence between the peaks in the absorption curveand the peaks in the TLF curve. This correspondence is shown in Fig. 27where the TLF spectrum (upper trace) is compared with the 2. 3 x 10~3 nmresolution absorption spectrum (lower trace). It appears that the TLFspectrum shows with much greater contrast the structure in absorptionwhich is superimposed on the continuum background absorption. The highercontrast in the TLF spectrum occurs because this measurement detects onlythe absorbed light that is emitted into the vx fluorescence channel while theabsorption spectrum includes all the light that is absorbed at the particularwavelength and either emitted into a broad continuum and many fluorescencepeaks or nonradiatively de-excited.

Time Dependence of the Re-emission From SO2: - The time dependenceof the v^shifted re-emission was examined using the 6 nsec resolutioncapability of the SDLF. In particular we were searching for a fast re-emission component characteristic of scattering. At very low gas pressures(< 1 mTorr) a significant fraction of the excited 803 molecules can diffuseout of the volume viewed by the spectrometer entrance slits before re-emission. Thus we do not expect to observe the intrinsic lifetime of the SOzre-emission. However, any anomolously fast component should have beendetectable. None was observed in experiments at a number of differentwavelengths. However, another interesting effect was observed. Super-imposed on the quasi-exponential decay following pulsed excitation at severalwavelengths is a sinusoidal component which oscillates at a frequency ofabout 100 MHz. This feature is clearly visible in Fig. 28. It is possible thatthis effect is real, rather than an experimental artifact, because it is notobserved at all excitation wavelengths, but it seems to be repeatable atcertain wavelengths. The oscillation may arise from an exchange of excitationenergy back and forth between the originally excited state and other states,perhaps in the triplet manifold responsible for the phosphorescence observednear 380 nm. A possibly analogous oscillatory exchange of energy occurs inthe system of two coupled oscillators. Presently we are attempting a moredetailed theoretical analysis of this possible effect in SO2, and simultaneouslyextending our experimental observations.

81

Page 87: study of resonance light scattering for remote optical probing

CO

03

S2co

oCOUJcco

150

100

50

0

ccoCOCD

ca

CO<«UJcc.o

30%

20%

10%

299.94

1299.92 299.96 300.00 300.04

ABSORPTION SPECTRUM WAVELENGTH (nm)

Figure 27 Comparison of TLF and absorption spectra, with variationsadjusted to the same amplitude. The wavelength scale of the TLF spec-trums has been shifted from its original calibration to match the absorptionspectrum obtained by Dr. R. J. Exton at NASA Langley Re'search Center.The TLF spectrum was obtained with a laser bandwidth of 0. 005 nm andmonochromator bandwidth of 0. 55 nm, centered on the Vj line fluorescence.The absorption^spectrum was obtained with a slit function resolution of0. 0023 nm.

82

Page 88: study of resonance light scattering for remote optical probing

CMOCO

COir>CMo

a•uCtf

0)CO

•s•i-l

OO

O<uenflin

T3<U

a.2COra

<u

•aa>

x)0]I

a>oC

I

<y MS cd

<D

00 tiCM M

O

83

Page 89: study of resonance light scattering for remote optical probing

As the SO2 pressure is increased from a few mTorr to 10 Torr, thefluorescence intensity integrated over time was observed to remain nearlyconstant. This result obtains because of the compensation between increasedabsorption and quenching discussed subsequently. However, the fluorescencedecay time, which is governed by the radiative and non-radiative decay ratesof the excited level from which the fluorescence originates, decreasessteadily as the pressure increases. For pressures of SO2 greater than about10 Torr the self quenching was observed to reduce the decay time to less thanthe 5 nsec laser pulse width.

Self-Quenching in SO2. - H. D. Mettee (ref. 20) in his article on theluminescence from SC>2 vapor has shown, by producing Stern-Volmer plotsfor low pressures, that the principal decay mechanism of an isolated SC>2molecule excited to the first singlet state is by fluorescence and that anyvibrational relaxation prior to emission is slight. In addition, Mettee con-cluded that electronic self quenching occurs at nearly every gas kineticcollision/with a cross section of • 188 nm2.

Mettee's experiments employed relatively broad band non-laser excita-tion. In our narrow band self-quenching experiments, the scattering cellwas pumped out to about 5 mTorr residual pressure through a stainless steelextension tube which runs from the scattering cell to a pumping and gashandling station. An Eck and Krebs valve of glass and Teflon polymerconnects the scattering cell to a flexible stainless steel bellows which is.connected to the extension tube. The flexible tubing allows the scatteringcell to be translated in and out of the laser beam without disconnecting anyof the tubing. The vacuum pump valve was closed and 10,8 Torr of anhydrousgrade SO2 was admitted to the scattering cell and connecting tubing. Thepressure was monitored with a Dow Corning 704 oil manometer. With themonochromator slits set on 50 |om, a 40 second count was taken at 310.71 nmon the vi fluorescent peak and at 311.71 nm on the continuum fluorescence.The laser wavelength was set at 299.96 nm and the power incident on thescattering cell was about 66 |jw. The laser power was measured with andwithout the scattering cell in the laser beam in order to determine the insertionloss of the cell and absorption of the SC>2 gas. The scattering cell was thenpumped down to four successively lower pressures and the above procedurerepeated. The absorption coefficient measured from the slope of the logplot from the data was 0.038 cm"1 Torr -1. This agrees with Exton's highresolution absorption measurements and P. Warneck's published results(ref. 53) for the absorption of SO2 at the 300.0 nm wavelength.

Figure 29 is a plot of our measurements of the fractional absorptiondivided by the self quenched fluorescence signal in the Vi channel for the 5different pressures. It is seen that the data is nearly linear with an inter-cept of zero; i. e. , it follows the usual Stern-Volmer dependence. In the

84

Page 90: study of resonance light scattering for remote optical probing

0.03i

P-PRESSURE OF S02

0.02

o>

0.01

0 4 6 8

P (TORR)

10 12

Figure 29 Stern-Volmer plot of SO^ self-quenching.

85

Page 91: study of resonance light scattering for remote optical probing

regime where the absorption is small the exponential factor l - e ' s isapproximated by <x£Ps and the detected signal Is is nearly independent ofpressure. In this case any increase in the fluorescence from an increase ingas density would be cancelled by a corresponding increase in self quenching.

Quenching of SO2 Fluorescence by Air. - Figure 30 shows a quenchedfluorescence spectrum from SO2 excited at 299.96 nm when 700 Torr of N2 isadded to one Torr of SO2. In this case the vx fluorescence peak was reducedin amplitude by a factor of about 190 when the buffer gas was added. The N2

Raman signal can also be barely seen in Fig. 30 along with two otherquenched fluorescence peaks shifted by the V2 and v3 vibrational quanta of SO2.Further out in the spectrum than shown in Fig. 30, in the 380-450 nm region,there are several phosphorescence bands whose intensities are almost asgreat as the partially quenched Vi fluorescence peak. Mettee has demon-strated that the phosphorescence levels are populated predominantly bycollisions. The phosphorescence lifetimes were observed to be severalmicroseconds long. When the room lights are turned out the faint bluephosphorescence emission from the laser beam in the scattering cell canalso be observed visually.

For the buffer gas quenching experiments the scattering cell was filledwith SOz gas at a pressure between 1 and 10 Torr. The Eck and Krebs valueat the cell was closed :and the fill line and gas handling system pumped downto about 5 mTorr. The vacuum pump valve was then shut off and either N2

from a laboratory bottle or air from the laboratory at about 100 Torr wasadmitted into the scattering cell and fill line. The N2 or air buffer gaspressure was monitored with a Hg manometer. The valve at the scatteringcell was cracked open and the pressure of the buffer gas was reset to 100Torr. It was found by monitoring the fluorescence signal and the absorptionof the laser beam through the scattering cell that the time required for thegas mixing was from between 5 and 10 min. After the buffer gas fill, thevalve at the scattering cell was shut off and 15 to 20 min. was allowed beforedata was taken. Using 500 pm input and output slits on the monochromator,a 40 sec. count was taken on and off of the Vj fluorescence peak as describedin the self quenching measurements. The power level of the laser wasmonitored and adjusted to keep the detected signal rate from saturating thephotomultiplier. These measurements were repeated for several more in-crements of buffer gas pressure up to 700 Torr. Several measurementswere taken with the N2 buffer gas at 900 Torr. After the buffer gas quenchingdata was taken the cell was pumped out and flushed several times with N2.The N2 Raman signal was then measured with 700 Torr of Nz in the cell.A background signal was then measured by shifting the monochromator1.0 nm off of the Vj fluorescence wavelength. The data was then reduced bycorrecting the count rate for single photon counts as described earlier andthen normalizing the count rates which were taken at different laser power

86

Page 92: study of resonance light scattering for remote optical probing

ccLUCO

c0)•a

• |H

g

CoX0)CO

CMOCO

0)oC(UoCOcot-.§

cvj

a:QCoI—oo

CVJOCO

toCVJ

OCO

O I-

CVJOCO

enazo

(51CO0)AH

oHOot-

o,O}05

o

a:cvj a> LO

o oC o0) .o oCO<y 5si

m

§

CCfl

JQ

87

Page 93: study of resonance light scattering for remote optical probing

1000

~ 800

OCUJ 600

OO

O

400

oO

200

TO 14,950 ATZERO AIR PRESSURE

I

I/, LINE INTENSITY

FOR S02 AT I TORRIN AIRX,=299.96 nm

AX =0.005 nm

KT25 cm2/Sr~

200 400 600 800

All? PRESSURE (TORR)

1000

Figure 31 Intensity of Vi-line fluorescence from SOj as a function of back-ground air pressure. The SO2 fluorescence is excited by incident light ofbandwidth 0. 005 nm at 299. 96 nm. The effective cross section for vt-fluorescence, measured by comparison to Na vibrational Raman scattering,is shown in the figure.

88

Page 94: study of resonance light scattering for remote optical probing

levels to a common power level. The background count rate, which was only0. 2 Hz was then subtracted from the data.

Figure 31 shows our results for quenching of the Vj^ fluorescence from 1Torr of SO% with the laboratory air as a buffer gas at five different pressures.

This figure demonstrates quite clearly that the variations of the SO2 Viline intensity is quite small over the normal range of variation of atmosphericpressures. Also shown here is the effective differential cross section forfluorescence into the v1 line by SO2 in 700 Torr for air pressure. This crosssection was determined by comparison against N2 vibrational Raman scatter-ing, using the previously measured absolute cross section for the latter at300 nm excitation. The result obtained is

a (SG>2,\4) '= 1 x 10"25 cm2/sr ± 35$

where the error estimate derives partly from the uncertainty in the N2 crosssection (±25$) and partly from statistical uncertainty in the SO2 and N2 countrates in the comparison measurement. It is significant that the SC^ crosssection is about four orders of magntidue stronger than that for N2 excited300 nm.

The results for air quenching of the SO2 v^ line are displayed in termsof a Stern-Volmer plot in Fig. 32. As in the case of SO2 self quenching, thelinearity of this plot suggests that SO2 absorption and subsequent fluorescencearises from a quasi-continuum. [See the discussion leading to Eq. (83)].

In order to determine the sensitivity of the Vj peak intensity to laserwavelength in the case of strong quenching, the spectrometer slits wereagain set to produce a wide (0. 5 nm) semi-rectangular slit function and thelaser was set to scan a small range. The resulting signal is shown in Fig.33. Although this signal is quite a bit noisier than the corresponding lowpressure result shown in Fig. 26, it is clear that there is some signalvariation. From this and other results we estimate the variation to be onthe order of ±30$ over the tuning range shown.

In another experiment the monochromator wavelength was shifted to311.71 nm, about 1 nm from the Vj peak, where the detected signal was fromthe continuum fluorescence. It was found that the continuum fluorescence at311. 71 nm also quenched at approximately the same rate as the Vj fluorescencepeak. It was also found that quenching of 1 Torr of SO2 with N2 from a labor-atory bottle of nitrogen gave the same results as quenching with the laboratoryair. In addition, using nitrogen pressures up to 900 Torr in the scatteringcell, we found that the quenching was still defined by the linear Stern-Volmerplot as in Fig. 32. No evidence of a deviation suggestive of scattering (i. e. ,an unquenched component of re -emission) was seen in any of these experiments.

89

Page 95: study of resonance light scattering for remote optical probing

300

K=0.27T'

200

0 200 400 600 800 1000

AIR PRESSURE (TORR)

Figure 32 Stern-Volmer plot of 803 quenching by air,derived from data shown in Fig. 31.

90

Page 96: study of resonance light scattering for remote optical probing

50

I TORR S02 ;700 TORR N,

FLUORESCENCE SIGNALAT 310.71 nm

299.94 nm

+0.050 +0.025 0 -0.025LASER WAVELENGTH nm

-0.050 -0.075

Figure 33 Tuned laser fluorescence spectrum ofVj-line in air.

A final point to be mentioned in this section is the polarization of the Vjfluorescence. The depolarization ratio p as defined previously was mea-sured using a uv»polarizer in the filter holder in front of the monochromatorcollecting lens. The depolarization ratio for the vt fluorescence at one Torrof SC>2 was 0. 71. When 700 Torr of N2 buffer gas was added to the scatter-ing cell the depolarization increased to unity. This increase in depolariza-tion is further evidence of the lack of a scattering signal (ref. 4).

91

Page 97: study of resonance light scattering for remote optical probing

IV. MODEL CALCULATIONS

In this chapter we consider several examples of applications of theresults obtained in this work.

Ground Based Lower Atmosphere Probe

The large quenched fluorescence cross section found for SOg opens thepossibility for concentration measurements of this gas at low concentrations.Initially, we consider two potential applications: measurement of sourcelevels and measurement of ambient levels in the lower atmosphere.

A typical source of SO2 is a smokestack whose effluent contains on theorder of 100 ppm of SO2 in nitrogen with other minor constituents. Supposethat the SOg concentration in the effluent is to be monitored from a distanceof 200 meters by a LIDAR system transmitting light at a wavelength of 300nm. The number of detected photons per joule of transmitted energy isgiven by

n =EXR2

where p is the number density of observed molecules, a is the differentialcross section (area/steradian) for the observed re -emitted light, L is thepath length from which re-emission is observed. A is the area over whichreturned light is collected, e is the optical efficiency, r\ is detector quantumefficiency, T is the two-way atmospheric transmission, E^ is the scatteredlight photon energy in joules and R is the range. We assume the followingsystem parameters:

(i

p = 2. 7 x 1015 cm3 (100 ppm at STP)L = 3 mA = ' 1m2

e = 10$ . • '•n = 20$ •T • = 50$

Ex = 6 xlO"19 J.R - 200m

Then, using the effective differential cross section found for the ^ line ofSO2 excited by 300 nm (nominal) radiation in air at STP,

a = 1.0 x 10~25 cm2/sr

92

Page 98: study of resonance light scattering for remote optical probing

we obtain n = 3 x 104 detected photons/ joule. This strong response wouldallow even our 100 |aW laboratory doubled dye laser to be used as a sourcemonitor. According to the results of this calculation, it would provide 300counts from 100 ppm of SC>2 in 100 seconds.

sources of the type assumed above can also be monitored remotelyby simple absorption (e. g. , against skylight) and by differential absorptionof Rayleigh/Mie scattering, the so-called DIAL technique (refs. 54,55). The1/e absorption length for 300 niri radiation in the assumed plume is about 4 meters.Thus, good contrast is available for these methods.

On the other hand, the 1/e absorption length for a typical ambient airlevel of 10 ppb SC>2 is about 40 km. In this situation, absorption techniquesare much more difficult to apply. However, the fluorescence signal appearsto be sufficiently strong to allow range -resolved measurements of SOg atthese levels at reasonable expense. Consider an experiment where 10 ppbof SO2 is observed by a LIDAR system with 100 meter resolution at 1 km.We assume that the experimental parameters are as described in theprevious example, with the following exceptions: p = 10 ppb (STP), R = 1 km,and L = 100 meters. Then from Eq. (88) we find

n = 4. 5 counts/ joule.

_This response is sufficiently strong to be observed in a practical instrument.We have pointed out previously that the decay time of the fluorescence re-emission from SO2 in air near STP is sufficiently short («5 nsec) to allowhigh resolution time-of-flight ranging. It would be desirable to investigatethe fluorescent return from 803 at slightly shorter wavelengths (say 270-280 run) in order to take full advantage of the dark daytime background atwavelengths shorter than 300 nm caused by absorption of sunlight in theatmospheric ozone layer. Also, the influence of water vapor and temperatureon the SC>2 quenching needs to be investigated. However, the present resultsare favorable indications that a LIDAR system observing SOg fluorescencecan be used to monitor sources and ambient distribution of this gas in thelower atmosphere. Such a probe might be applied, for example, to map theSOz distribution far downwind of a source.

Ground-Based Upper Atmosphere Probe

Sodium and potassium layers at altitudes near 80-100 km have alreadybeen probed successfully from the ground using dye lasers and fluorescencetechniques (refs. 15, 16). In this section we consider the possibility of aground-based, spatially resolved probe of the O3 layer at an altitude of 10-35 km. .This application is important because the O3 layer forms the primaryshield of solar middle UV from the ground, and because it is an importantindicator of chemical chains and physical processes in the stratosphere.

93

Page 99: study of resonance light scattering for remote optical probing

The ozone layer provides many orders of magnitude absorption atwavelengths near 250 nm, falling off gradually toward 300 run. In order toachieve range resolution we will assume that a transmission wavelength ischosen near 300 nm, such that there is approximately 25% absorptionthrough this layer. The number n of detected photons per joule of trans-mitted energy will be estimated from a slightly different procedure thanused previously. This number can be calculated from the expression

n~(0 .25) (89)EXR2

where in addition to the parameters defined previously, the factor (0.25) isthe assumed absorption, £ is the efficiency of re-emission into the observedwavelength range following absorption of a photon, and the factor 4ir arisesfrom the (approximately) isotropic re-emission into 4n steradians. It isinteresting to note that for SC>2 in air at STP, our previous results yield§ — 10~6, determined primarily by quenching. At an altitude of 25 km,where collision frequencies are down by a factor of about 25 from STP(ref. 54), 5 should be on the order of 2. 5 x 10~5 for SO2. In the case of O3,excitation into the absorption bands near 300 nm is dissociative. Thus.there-emission from O3 may be similar to that of the resonance scatteringfrom I2 observed by Holzer et al. (ref. 4) in the visible. A rough estimateof in this case is given by the following product:

_ /fraction of radiation in \ v, /mean time to radiate \\observed spectral linesy>^\ \jnean time to dissociate^

do-1) u sec

\10 13sec/

Here the mean time to radiate is calculated from the integrated absorptioncoefficient for the Hartley band, estimated from Fig. 9, using a well knownrelationship (ref. 12) between these quantities. The mean time to dissociateis set equal to the time required for an oxygen atom with 2 ev of kinetic energyto travel 0. 5 nm.

In order to calculate the number of detected photons per joule asgiven by Eq. (89), we assume the following system parameters:

Ex = 6 x 10"19J T = 10%5 = 10~6 A = 10 m2

e = 10% R = 25 kmr\ = 20%

These assumptions yield

r\ = 1. 0 detected photons/joule

94

Page 100: study of resonance light scattering for remote optical probing

which is well within the range of experimental feasibility. The experimentaloutlook is simplified by the fact that a narrow band, precisely-tuned sourceshould not be needed because the relevant O3 absorption (Hartley band) is abroad continuum. However, when the sun is near the zenith, substantial daylightbackground will be encountered. See ref. 56 for an analysis of the effect ofthis background. We should note that the assumption of 10% two way trans-mission of 300 nm radiation to 25 km altitude is probably optimistic for sealevel operation, but should be achievable from elevated ground (say 5>3 km).The requirements for high altitude and large collector area are similar tothose for astronomical observatories. Thus the possibility of involving anobservatory in LIDAR studies of the upper atmosphere is suggested.

Airborne Measurements

A LIDAR system carried by a high flying airplane can be used to deter-mine the concentrations of many high altitude species. In this case powerand collector area limitations can be compensated at least partially byproximity. As one example we consider a probe of ozone at 25 km by aplane flying at an altitude of 15 km. We assume a i m 2 collector, T = 50$and other system parameters as in the previous example. Then

n = 3. 3detected photons/joule.

Likewise, an airborne system can be used to probe many other highaltitude species. For example, we consider the system described above,but observing a 100 meter column at a range of 300 meters. Suppose that1 joule is transmitted and that the expected number of detected photons isrequired to satisfy the condition

n 2 100

in order to obtain reasonable statistical accuracy. Then from Eq. 88 weobtain as a condition for a successful measurement

pa = 6 x 10'14 cm'1 sr'1 (90)

In Table HI we show estimated effective differential cross sections forfluorescence from a number of molecules and atoms at 20 km altitude,and the corresponding minimum densities which can be measured fromEq. 90.In the case of NO, feasible measurements must await development of doublingtechniques or primary laser sources at 227 nm, but a recent publication(ref. 57) suggests that this development is likely to occur in the near future.

95

Page 101: study of resonance light scattering for remote optical probing

TABLE III

Estimated effective differential cross sections for molecular fluorescence(scattering in the case of O3) in the atmosphere at 20 km altitude, and con-sequent minimum measurable concentrations (mmc) for the assumed airborneLIDAR system calculated using Eq. 90. The cross section for SC>2 was takenfrom experimental results obtained in this work. That for NC>2 was calculatedfrom data presented by Fouche, Herzenberg and Chang (ref. 11). The NOcross section is obtained from absorption data of Bethke (ref. 21), assuminga 10~6 sec lifetime for radiative decay and quenching at every gas kineticcollision. The OH cross section was calculated similarly using absorptiondata presented by Penner (ref. 58). The scattering cross section for O3

(excited by absorption in a dissociative continuum) is calculated by multi-plying the absorption cross section from Fig. 9 by the re-emission efficiencyestimated in this Chapter. The mmc are expressed as absolute numberdensities and also as parts per billion (ppb) referenced to a total moleculardensity of 1. 85 x 1018 molecules/cm3 at 20 km (ref. 59). Estimate re-emissionefficiencies § are calculated such that (4n/§) times the fluorescence crosssection equals the corresponding absorption cross section. Also shown arecorresponding results for the atoms listed in Table IV. The atomic fluorescencecross sections were calculated for the center of Doppler-broadened lines(220°K) using oscillator strengths collected in ref. 60. (The effects ofcollision broadening and hyperfine structure in the absorption line has beenneglected in these calculations.)

EstimatedRe-emission

Molecule Efficiency g

SO2 1.6 x 10"5

(300 nrn)

NO2 10(400-500 nm)

-7

O3(260 nm) 10"6

NO(227 nm) 3 x 10"3

OH(308 nm) 1.6 x 10~4

Atoms in 0. 001 to 0. 1Table IV

EffectiveCross Section

2 xlO"24cm2/sr

2 x 10"27

1 x 10-24

3 x 10~22

2 x 10-22

10-18 to 10-14

cm2/sr

mmcmmc in

ppb at 20 km

3 x 1010 16molecules/cm3

3 x 1013

6 x 1010

2 x 10s

3 x 108

6 to 6 x 104

atoms/cm3 (!)

16000

32

.11

.16

96

Page 102: study of resonance light scattering for remote optical probing

In addition to the molecules listed in Table III, there are many otherspecies which can exhibit moderate to strong re -emission at accessiblevisible and uv wavelengths. Among these^ are the aromatic hydrocarbons,aldehydes, and various radicals such as CN and 03.

In addition to its use to estimate minimum measurable concentrations,Eq. (90) can be used to demonstrate that under many conditions the LIDARtechnique used with fluorescence will be more sensitive than absorptiontechniques. First, we note that a relationship between the effectivedifferential fluorescence cross section and the corresponding absorptioncross section is given by

o ~ (?/4TT) aabs (91)

where the re-emission efficiency §has been defined following Eq. (89).Thus the condition for a feasible observation (with the experimental parameters assumed above) becomes

But the condition for a feasible observation in an absorption measurement(for example, using the DIAL technique) is

H 10

where H is the thickness of the observed layer. Thus the assumed LIDARsystem will be more sensitive than a system using absorption measure-ments if

|r < 1.33 x 106 km (92)

Note that if the DIAL technique is used to obtain range resolution in anabsorption measurement, H corresponds to the range resolution. Theresults indicated in Table III suggest that at 20 km altitude Condition (92)is likely to be satisfied for NO, OH and atoms, but not necessarily forSO2 and O3.

97

Page 103: study of resonance light scattering for remote optical probing

Measurement of Upper Atmosphere Constituentsfrom a Satellite

Satellite LIDAR measurements are more difficult because of powerand collector area limitations, as well as the large working distances re-quired. On the other hand, transmission is high and many effectivefluorescence cross sections are much larger than in the lower atmospherebecause of reduced quenching. (The collision frequency is lower by afactor of 4 x 106 at 100 km altitude than at sea level - ref. 59. ) The generalnet effect is that measurements are more difficult from a satellite. How-ever, we will show that there are LIDAR type measurements which appearquite feasible.

In order to develop quantitative estimates, we assume as in theprevious section that a 1 joule pulse is transmitted, that returned light iscollected over a 1 m2 area, and that the LIDAR system observes a 1 kmcolumn at a range R. Other experimental parameters are taken to be thoseassumed in the previous example. Then, as a condition for a successfulmeasurement, we obtain from Eq. (80)

(93)

When the number densities of various species at high altitudes are takeninto account, Eq. (93) implies that a successful measurement requires alarge cross section. However there are many atoms with fluorescencecross sections on the order of 10~13 to 10"16 cm2/sr, for which very smallconcentrations can be measured. A representative group of these atomsis shown in Table IV. In Fig. 34 we show resulting estimated minimummeasurable concentrations as a function of altitude, calculated from Eq. (93)assuming a satellite altitude of 250 km. The total density as a function ofaltitude is shown on the figure for comparison.

Also shown are the estimated minimum measurable concentrations forseveral molecules. It is immediately apparent that the situation formolecules is much less favorable than for the atoms discussed previously.This result obtains because of the multiplicity of available states introducedby the additional degrees of freedom of a molecule. This multiplicityreduces the re-emission cross section in three ways: first, because onlya fraction of the molecules will be in initial states leading to the observedtransition; second, because the strength of the transition is divided amongall accessible excited states; and third, because the re-emission isdistributed among all allowed downward transitions. Thus fluorescencecross sections for molecules are smaller. Estimated values used inpreparing Fig. 34 are shown in Table V.

98

Page 104: study of resonance light scattering for remote optical probing

ATOMSLISTED

INTABLE

L06|Q NUMBER DENSITY

Figure 34 Minimum measurable concentrations of atoms and OH radicalunder conditions assumed for satellite probe. The total molecule densityas a function of altitude, from ref. 59, is also shown for comparison.

99

Page 105: study of resonance light scattering for remote optical probing

TABLE IV

Atoms for which strong fluorescence under excitation on resonance linesbetween 250 and 900 run has been demonstrated or is predicted on the basisof available transitions with large oscillator strengths. Unquenchedfluorescence cross sections for these atoms calculated for the center of aDoppler-broadened line (200°K) are on the order of 10~13 to 10~16 cm2/sr(hyperfine structure effects neglected). Corresponding excited state life-times are on the order of 10"9 to 10~6 seconds. Excitation wavelengths forthese atoms, and corresponding oscillator strengths are collected inref. 60. Atoms indicated by an asterisk (*) produce strong fluorescence atwavelengths shifted from that of the incident light, allowing discriminationagainst Rayleigh and Mie scattering.

Approximate Excitation Approximate ExcitationAtom Wavelength Atom Wavelength

Li 671 nm Mn 403 nmNa 589 Re 346K 770 *Fe 300Rb 795 *Ru 299Cs 894 *Os 353

*Cu 327 *Co 341Ag 328 *Rh 343Au 268 *Ir 250Mg 285 '*Ni 337Ca 423 *Pi 276Sr 461 *Pt 266Ba 554 *A1 394Cd 326 *Ga 403Hg 254 *In 410

*Rare Earths 400-900 *T1 378*Sc 391 *Sn 286*Y 408 *Pb 283Cr 427 *Bi 307MO 390 *C 286

100

Page 106: study of resonance light scattering for remote optical probing

TABLE V

Estimated effective differential cross sections for molecular fluorescence(scattering in the case of O3) in the atmosphere at 100 km altitude. Thesecross sections were calculated from measured absorption coefficients(refs. 19,20,21,24,59), using the equation Ofi - aabs (5/4rr). Re-emissionefficiencies § were estimated assuming no quenching since the collisionfrequency at 100 km is small (2 x 103 sec"1). For SC>2 and NO2, | takes intoaccount the effects of the anomalously long fluorescence decay times (ref. 39)which (are presumed to be related to non-radiative decay. The value of § forO3 is diminished by dissociation; we use the value estimated earlier in

this report.

Molecule

S02

N02

03

NO

OH

Estimated Re-emissionEfficiency £

1 x 10~3

5 x 10~3

10~6

lo-1

10-1

Cross Section

1 x 10~22 cm2

1 x 10"22

1 x ID'24

1 x 10-20

1.3 x 10'19

/sr

Equation (93) can be used to estimate in what cases a LIDAR measure-ment will be more sensitive than an absorption measurement, as was donepreviously for the airborne system. In the present case, we find that theassumed LIDAR system should be more sensitive when

, 1 .33xl0 3 km (94)§ \ R /

where R is the observation range. This condition will be satisfied, forexample, if H = 20 km and R = 150 km for 5 £ .034. All atomic species anda few of the molecules considered will satisfy this condition at altitudes> 100 km.

101

Page 107: study of resonance light scattering for remote optical probing

V. CONCLUSIONS

Major tasks supported by this contract are:

1. A classical theoretical study of the amplitude and time dependenceof re-emission from an isolated molecule following pulse light excitation, asa function of separation from resonance. (Chap. n). This study includesformulation of the problem, analysis of limiting cases and numerical solu-tions for intermediate cases in the transition from scattering to fluorescence.

2. A quantum-mechanical study of the problem described above. (Chap.II and Appendix C). The resulting expressions for the time dependence ofthe re-emission were shown to agree with the classical results in significantlimiting cases. This agreement was taken as an indication that the classicalresults are essentially correct.

3. Theoretical analysis of the depolarization of scattering followingexcitation into a dissociative continuum, in which a quantum interferenceeffect that tends to reduce depolarization was discovered (Chap. II andAppendix D).

4. Qualitative estimates suggesting the possibility of very strongscattering from O3 excited in its Hartley band, which is a dissociativecontinuum (Chap. IV).

5. Measurement of the absolute NO2 cross section for line fluorescenceexcited in air near STP by visible light (488 nm) (Chap. III).

6. Measurement of the O3 vibrational Raman cross section for visiblelight (488. 0 and 514. 5 nm) excitation (Chap. III).

7. Assembly of a doubled dye laser spectroscopy system from com-ponents on hand (Chap. III).

8. Measurement of absolute N2 and O2 vibrational Raman cross sectionsfor excitation at 300 nm, using the doubled dye laser system (Chap. III).

9. Detailed quantitative study of SO2 Vj-line fluorescence and quenchingof this fluorescence by air (Chap. in).

From the consequent results and others described in this report, wehave drawn the following conclusions:

Near Resonance Scattering

Analysis of the time dependence of re-emission from an isolated moleculeexcited by a light pulse shows a clear transition from delayed re-emission

102 , . *

Page 108: study of resonance light scattering for remote optical probing

(fluorescence) to "instantaneous" re-emission (scattering) as the separationfrom resonance is increased beyond substantial overlap between spectraldistributions of the absorption line and incident light. This result was usedto conclude tentatively that in a real gas the re-emission would be both in-stantaneous and insensitive to such effects of molecular interactions asquenching, collisional depolarization, and collisional relaxation when theseparation from resonance is several times larger than:

A. The collision linewidth, and

B. The range of significant inhomogeneous broadening (statisticalaverage of line shifts produced by Doppler effect and molecularinteractions).

However, we cannot regard this conclusion as firmly established at thepresent time. Early results from an alternative quantum electrodynamicanalysis suggest that the re-emission intensity in primary lines (i. e., linesfrom the originally excited intermediate state) will be independent of pressureunder the conditions specified above, but that collisional depolarization andrelative intensity of collision lines (satellite lines produced by energy ex-changed during collisions) may remain equal to their values in fluorescenceout to much greater separation from resonance. Our inability to draw firmgeneral conclusions about the character of re-emission as a function ofseparation from resonance in a real gas is related to the location of thephenomena of interest in the far wings of an absorption line, where presentlydeveloped theories falter. Tuned laser fluorescence studies are expected toproduce both incentive and information for further developments in this area.As a case in point, in work supported by GE we investigated collision-to-primary line intensity ratios as a function of separation from resonance inlow pressure (200mTorr) 12 vapor. The incident light was from argon andkrypton lasers that were single-moded and tuned over various gain curvesusing a tilted etalon. Over substantial tuning ranges (several times theDoppler full width at half maximum in one case) this ratio could be observedprecisely and was found to remain essentially constant. This result supportsthe suggestions of the alternative quantum electrodynamic theory. However,it can also be explained on the basis of the original conclusion as arisingfrom effects of inhomogeneous broadening produced by line shifts.

Scattering Following Excitation in Dissociative Continuum

In addition to near-resonance enhancement, we considered also the typeof scattering that results from excitation in a dissociative continuum. Holzeret al. showed that this type of scattering from I2 excited by blue light (488 nm)is substantially stronger than nitrogen Raman scattering. Our rough calcu-lation in Chap. IV for the intensity of Raman scattering from O3, excited inis dissociative Hartley band between 220 nm and 310 nm, indicates that inthis case the scattering should be very strong, with a cross section on the

103

Page 109: study of resonance light scattering for remote optical probing

order of 10~24 cm2/sr, or about five orders of magnitude stronger than thenitrogen Raman scattering.

Fluorescence

At the outset of this work it was presumed that the advantages of scat-tering over fluorescence as a gas probe would be important in atmosphericmeasurements. Accordingly, the enhancement of scattering in the approachto resonance and the ultimate shift from scattering-like to fluorescence-likere-emission in this approach received strong initial attention. However,during this work it became increasingly apparent that the greater strengthof fluorescence is crucial to many measurements of low concentrations inthe atmosphere from substantial distances, and furthermore, the potentiallydetrimental characteristics of fluorescence as a gas probe (e. g., quenching,time delay at low pressure) are often not serious impediments to these mea-surements. Specifically, fluorescence measurements with a LIDAR systemshould be useful under the following conditions:

1. The species whose measurement is sought exists at low concentrationin a carrier gas that dominates line broadening and quenching.

2. The pressure of the carrier gas is known with sufficient accuracy tomake line broadening and quenching corrections, and it is high enough toshorten the re-emission delay of observed fluorescence to the point wheretime-of-flight ranging can be used.

3. The quenched fluorescence of the target species is much strongerthan ordinary RS and has a characteristic spectral distribution that allowsits identification.

With respect to the time dependence, we and others have observed thatfluorescence decay times are sharply reduced by quenching to times on theorder of a few nanoseconds in air near STP. Even at the low pressures andcollision frequencies encountered at high altitudes, the consequent near11 natural" decay times of fluorescence for many molecules are not so longas to preclude useful range resolution. For example, an exponential decaytime of 1 microsecond allows a resolution element of about 150 meters.With respect to the strength of fluorescence it has been found that even underrapid quenching in air near STP, characteristic fluorescence can be manyorders of magnitude stronger than ordinary Raman scattering. For quanti-tative measurements in a system such as the lower atmosphere, dominantquenching rates must be determined accurately. In many cases, where theabsorption and/or fluorescence cross sections have spectral structure on afine scale, the spectral distribution of the exciting light must be controlledprecisely. Furthermore, the sensitivity of these cross sections to tempera-ture and minor species such as H2O and CO2 must be investigated. Never-theless, quantitative experimental results such as those we have obtained

104

Page 110: study of resonance light scattering for remote optical probing

for SO2 (see below) suggest that fluorescence measurements of minor speciesin the lower atmosphere may be quite worthwhile.

Experimental Results

In the experimental part of our work, re-emission from NO2, I2 and O3

was observed in the visible, using an argon laser source, and re-emissionfrom N2, O2 and SO2 was observed in the ultraviolet near 300 nm. We foundthat the characteristic Vj line fluorescence from NO2 in air near STP isslightly more than 100 times enhanced over the vibrational Raman scatteringfrom N2, in reasonable agreement with results calculated from data publishedby Fouche, Herzenberg, and Chang (ref. 11). This enhancement is probablynot sufficient to allow practical measurements of ambient levels of NO2 in theatmosphere, but there is a possibility that much stronger enhancement willbe found under excitation at particular wavelengths in the blue, whereSackett and Yardley (refs. 35, 36, 37) have observed fluorescence componentswith anomalously fast decay.

Perhaps our most useful observation with respect to atmospheric probesis the very strong fluorescence from SO2 in air near STP excited by light near300 nm. The intensity of this fluorescence was measured to be 10* timesstronger per molecule than vibrational Raman scattering from N2, which it-self was found to be several times larger than a (1 Ascatter)4 extrapolationfrom measurements in the visible.

Model Calculations

Effective cross sections for fluorescence from OH, NO and a largegroup of atoms were estimated from available data in Chapter IV. Thesecross sections correspond to excitation in the range from 250-900 nm,except for NO which must be excited near 227 nm. The-estimated moleculecross sections, though smaller than those for atoms by, typically, 6-8 ordersof magnitude, are still sufficiently large to allow useful measurements.

Measured cross sections for v^line fluorescence from SO2, excited near300 nm, and estimated resonance Raman scattering cross sections for O3

excited between 250 and 300 nm were used as a basis for model calculations.These calculations indicate that the cross sections are sufficiently large toallow range-resolved LIDAR measurements of source levels of SO2 and evenambient levels (on the order of 10 ppb) of SO2 and O3 in the lower atmosphere.Further, it appears likely that the O3 cross section is sufficiently large toallow range-resolved LIDAR measurements of the stratospheric O3 layerbetween 10 and 30 km altitude from a high altitude ground-based station oran airborne station.

105

Page 111: study of resonance light scattering for remote optical probing

The large fluorescence cross sections predicted for numerous atomicspecies at accessible excitation wavelengths allow concentrations of a fewatoms/centimeter3 to be measured at close range (300 meters) by an airborneLIDAR probe. Some molecule concentrations (e. g., SO2 and O3) may be mea-surable also. The large atom cross sections establish a good possibility thatmany atom concentrations in the upper atmosphere can be measured from asatellite LIDAR probe. However, the possibility for satellite LIDAR mea-surements of existing levels of molecules and radicals such as OH appearsto be marginal at best.

*

106

Page 112: study of resonance light scattering for remote optical probing

APPENDIX A

Distinction Between Fluorescence and Resonance Raman Scattering

Various criteria have been used to distinguish between Raman scatteringand fluorescence excited near or on resonance. The original distinction isthat the frequency shift in Raman scattering is independent of incident wave-length (frequency), whereas in fluorescence the frequency itself, rather thanthe shift, is constant. Thus as incident light frequency is shifted, the Ramanscattering frequency will shift comensurately, while "fluorescence appearsonly at wavelengths characteristic of molecular transitions. However, thisdistinction is difficult to apply in the case of excitation into a true or apparentcontinuum (e.g. , NO2 or SC>2 absorption) where fluorescence would also beexpected to shift with incident wavelength. It is also difficult to apply,although presumably operative, in cases which involve excitation very nearresonance, because of problems associated with high resolution spectroscopyat low light levels.

Hibben(ref. 61) expresses a common distinction that scattering is in-stantaneous whereas fluorescence shows an exponential decay. However,strongly quenched fluorescence in high pressure gases, and fluorescencefrom many solids (e. g. dyes) is also very fast (10~10 to 10~12 sec.) such thatits decay time is difficult to sense. It is possible to avoid this ambiguity bydefining scattering as an intrinsically'instantaneous process; i. e. , a processwhich is effectively instantaneous in an isolated molecule. Yet this defini-tion is obviously inapplicable to fluorescence from solids and liquids, andnot even be generally meaningful for gases because of the expected sensitivityof re-emission on the borderline to the line broadening produced by molecularinteractions.

The relevant work of Behringer and associates, extending from thepresent back to at least 1956, pertains mostly to "pre-resonance" andresonance excitation in solids and liquids. In a fairly recent review (ref. 25),Behringer appears to identify resonance scattering as that component of there-emission excited near resonance with spectral characteristics similar to"ordinary" scattering excited well off resonance. On the other hand, featuressuch as an underlying continuum which appears in the approach to resonanceare identified as fluorescence. Behringer notes that the relationship betweenthese phenomena is incompletely understood.

In 1970, Holzer, Murphy and Bernstein (ref. 4) published an importantpaper in which they identified a new type of resonance scattering, whichresults from excitation in a dissociative continuum. In this paper theyspecified a consistent set of distinctions between Raman scattering and fluo-rescence. In particular they noted that resonance Raman scattering is

107

Page 113: study of resonance light scattering for remote optical probing

usually strongly polarized and always insenstive to quenching, in distinctionto fluorescence. Subsequently, Berjot, Jacon and Bernard (ref. 9) identifiedas fluorescence those processes which proceed predominantly through asingle intermediate level, whereas scattering results when transitionsthrough at least several intermediate levels are significant. This identifi-cation has the advantage of associating distinctive selection rules with thefluorescence, with the result that fluorescence and scattering can usually bedistinguished by the spectral character of the re-emission.

Our own distinction is strictly utilitarian. We define scattering to bere-emission which is effectively instantaneous and insensitive to effects ofmolecular interactions such as quenching, collisional depolarization and thegross spectral broadening produced by collisional transfer. However, neitherwe nor others to our knowledge have yet demonstrated unequivocably whetheror not re-emission excited off resonance, but predominantly through a singlelevel, satisfies these criteria.

108

Page 114: study of resonance light scattering for remote optical probing

APPENDIX B

Quantum Mechanical Interferences

The quantum mechanical expression for the scattering cross sectionis given by

12) = ) Fn ) ^2 6(u)rf

c4ft i-> i->n

(x)_ (D2)fr rn

ID ,.rf- iy /2

r

which is identical to Eq. (38) except that the average over the fractionalpopulation of initial states Fn and sum over final states is indicated explicitlyand the 6-function ensures energy conservation. Within the absolute square,the sum over intermediate states {r} may include terms of opposite sign(or in the general case, complex terms with different phases) such thatthey tend to cancel. Since the relative magnitude of these terms depends onthe incident frequency, interference effects can produce oscillations in thecross section as incident frequency varies. (Notice that contributionsstarting from different initial states and/or ending on different final statesdo not interfere, because the absolute square is performed for each term inthe sum over initial and final states. ) The most obvious situation whereinterference occurs is when incident frequency increases through a resonance,changing the sign of one of the terms. Interference effects of this type havebeen predicted in ref. 18. They are responsible for the dimunition ofdepolarization in scattering excited within a dissociative continuumwhich we have predicted and observed, and they also produce theobservable effect in level crossing spectroscopy, where levels areshifted by an applied field.

From another standpoint, the possibility of this type of interferencemakes accurate estimates of scattering enhancement from oscillator strengthsuncertain in cases where more than one intermediate level is likely tocontribute significantly. A case in point is our estimate of the NO crosssection at 250 nm, where we simply summed contributions from variousvibrational bands algebraically, ignoring possible interference.

109

Page 115: study of resonance light scattering for remote optical probing

This procedure appears to be a good approximation when the separationfrom resonance is large compared with the separation between intermediatelevels which contribute significantly to the scattering. To demonstrate thispoint, we consider the case where the relevant levels are associated withdifferent rotational angular momentum, but the same vibrational andelectronic states. Employing the Born-Oppenheimer approximation, weseparate the rotational eigenfunctions R and the electronic-vibrationaleiginfunctions T to obtain for one of the terms in the absolute square

(Dz)T'R', T»R» (D2)T»R», TR

Here unprimed, single-primed and double-primed letters designate initial,final and intermediate states, respectively. Now if the separation fromresonance is sufficiently large, WRHR may be neglected and the sum over R"performed through closure (ref. 28). Then, if the absolute square is com-pleted and the sum over all final states (including scattered radiation states)is performed, one obtains simply the absorption cross section. Workingbackward, the relative magnitudes for various components in the sum yieldthe branching ratios for various types of scattering. This argument may beextended to contributions through different vibrational and electronic stateswhen the separation from resonance is sufficient. Unfortunately, in manycases, such as our estimate of the NO scattering excited at 250 nm, it isnot. In such cases, our estimation procedure yields an "expectation value"but with a large uncertainty. A better estimate appears to require a muchmore detailed calculation.

110

Page 116: study of resonance light scattering for remote optical probing

APPENDIX C*

Quantum Calculation of Time Dependence of Re-emission

We desire a simple quantum mechanical model and formulation, whichdescribe the time dependence of the scattering of a coherent pulse of lightthat is emitted from a laser pump source at a well-defined time and scattersfrom a quantum scatterer located at the origin. Our treatment is essentiallythat of a fixed single scatterer at the origin which would seem to preclude abinitio many-particle effects which give rise to inhomogeneous broadeningof spectral lines both from molecular collisions and thermal motion of thescatterers, and collisional-nonradiative de-excitation (quenching) of excitedstates. The inhomogeneous broadening can be taken into account by suitablyaveraging over initial states of the scatterer, while the quenching can betaken into account, at least heuristically, by a suitable collisional renormal-ization of the intermediate state propagators.

For the internal states of the scatterer, we will consider the simplifiedmodel of a three-level quantum system with energy levels 0, 1, 2. We willassume that there are allowed dipole transitions between CCl, 1^2. Thestates 0 and 2 can be chosen as the ground state and low-lying excited state,which could be contained within the ground state manifold, e. g. , a vibrationalstate. The pulse of light which is introduced into our system will have acentral frequency U>L and a frequency breadth A; the time breadth of the co-herent pulse ~1/A. We can view this problem in the following qualitativeway. If the packet is far removed from the origin, there is no interactionbetween it and the scatterer; the interaction commences when the leadingedge of the wave packet overlaps the scatterer. We can crudely break thescattered light emission up into two time regions: first, the emission ratewhen the wave packet spatially overlaps the scatterer; and, second, afluorescence which occurs at a time after the packet has passed the origin.Usually one calculates fluorescent emission by initiating the system in anunstable state at time t - 0 and then allowing the system to radiatively decay.Here our initial condition is a wave packet of light and a scatterer in theground state. For near-resonance cases, the wave packet will have Fouriercomponents which mix the ground state with an excited state, and hence therewill be a certain probability that the scatterer will be left in the excited stateafter the pulse has passed. This condition will then serve as the initiationof the unstable state from which the system can subsequently fluoresce. Therelative and absolute strengths of the "scattering" and "fluorescence" sodefined will depend upon the parameters of the wave packet and in particularon the nearness to resonance of the central pulse frequency tw^.

-This appendix is taken from GE Report 73CRD196, authored by Seth.D.Silverstein. The work was supported by Contract NAS1-11624 from NASALangley Research Center.

Ill

Page 117: study of resonance light scattering for remote optical probing

Before we go into the specific quantum mechanical representation of thepulse and the details of the calculation, we will first outline the generalformulation of the scattering problem with propagating wave packets (ref. 62).

Let us consider first the state at time t = -t0 where the center of thewave packet is sufficiently removed from the origin that there is no overlapand hence no interaction. If we denote K as the sum of the free scatter andelectromagnetic Hamiltonians, the wave function at time t - -to is given by

_iKto i MT \ i i \

Here ] Yj c> represents the wave function for the light pulse, and | $0 )ground state of the scatterer. We note that in standard time -dependent per-turbation theory the initial states are usually taken as eigenstates of K andhence are stationary states in the absence of interaction. This is not ap-propriate for the case considered here because we desire that the pulse prop-agates in space and hence is nonstationary. Let us now turn on the inter-action, so that the wave function at time +t is given by

| Y i c > | $ 0 > f (C2)

where

K = K + V

We desire to obtain the probability amplitude for the system being in a non-interacting final state at a time t,

|Y f( t)> = e ~ ' > l $ ' > • <C3)

This is given by

< Y ' | <$'|eiKte"i:K1[t + t o )e i K t o |Y i c>|$ 0> . (C4)

If we define

U(t + t0) = eL e "^-o/e^i0 , (C5)

U possesses the standard perturbation expansion,

•t ^t

. (C6)

112

Page 118: study of resonance light scattering for remote optical probing

We are interested in a scattered photon; hence the terms of interest will bederived primarily from the second-order terms in the expansion of U. If thetime of transit from the laser source to the scatter at the origin is longcompared to the time width of the pulse, we can always set the initial time- 10 -• - », with the condition that the wave packet is centered at the originat t = 0.

We now want to consider the quantum mechanical representation of theinitial pulse state |^ic). We choose \^ic} in a form such that the expectationvalue of the electric field operator describes a pulse propagating in the +zdirection;

1C ~ ~

l

e Re/~x /

•rn

duu E(UJ - u.'-iuut

<C7)

Here ^ is a unit polarization vector in the x direction. The electric fieldoperator is given in terms of photon creation and annihilation operators by

-i(k- r -lOb-t)~ ^ K _ i (k 'r-u) . t)

k, X

(C8)

We can achieve a semi-classical representation of the form(C7)by requiringthat |Yic) be eigenstates of the destruction operators,

, . .k, X ' ic

a, . IT. > for k, X e ( k >0, k, =0, X =X )k, X ic ~ z J- x

0 otherwise.(C9)

The "coherent state" representation of the packet wave function (ref. 63)satisfies this condition:

I V = exp Iu >0

(Vk.X.-l°k.l(CIO)

Here |0> is the photon vacuum state and the wave function lYic) is normalized.We can readily verify that (C9) satisfies (CIO) by expanding the exponential,operator in a power series and then operating on the result with a , ^ , e.g.,

113

Page 119: study of resonance light scattering for remote optical probing

. + m °° , vm-1(akz.Xx; ... . V" kzXx kg; inN

^ - ' } k p k z *•** kz2- (m-1)! - I0>

m=0 m=l

= a. 6, , 6, . expfa* . a, 1|0> .k k, k X, Xx ^L k, Xx k J1

Z. t^t t-t ~ . ^ Z

We previously commented that the quantum state representing the propagatingwave packet is nonstationary and hence not an eigenstate of the free electro-magnetic field Hamiltonian. This is apparent from the form of (C7)exponential form of the operators exhibits an undetermined photon number.As is well known, photon number and phase are canonical operators, and itis this uncertainty in photon number which gives the desired aspect of thecoherent propagation of the pulse. We note that our model of a single scattererin vacuum rules out dispersive distortion of the pulse and such effects willbe neglected.

We now want to specify the normalization for the wave packet function.We use a similar procedure as given in the preceding classical calculation,appropriately modified for the quantum case,

^f4rr /J —(

dt I <Y. |E(r = 0 , t )xK(r=0 , t ) U. >| Y=Nftuu T + const. . (C12)' m ' ~ ~ ' i n 1 i-i. I v * * | \ * » | * - J \ * V f i . / -^^V \* ^» "I \ * •

TtTT I 1C *^^ 1C

That is, we want to describe the pulse in terms of an effective photon numberN. The constant term is the vacuum contribution and does not depend uponthe pulse. Upon insertion of the appropriate forms for the electric andmagnetic field operators, we- obtain

. (C13)

Taking a Lorentzian form of a(u)),

1(C14)

our normalization condition [Eqs. (41), (42)] gives

114

Page 120: study of resonance light scattering for remote optical probing

We note that we have added a factor of (UUL/UU) to a standard Lorentzian inEq, (Cll). This is done for convenience only and can be replaced by unity toterms of order A/I«L. We note that our analysis will neglect terms of orderA/WL relative to unity as they will be very small for optical frequencies ofinterest.

For a Gaussian wave packet we take

/TtrT~a(uu) = A '\/—^— exp[-(u)-uuL)3 /2A3] . (C16)

to terms of order exp[-u)L/A2] relative to unity, we have

A 3 2/rfcVNAg = I^A ' (C17)

In our calculation of the time dependence of the scattering in the approach toresonance, we will work with the Lorentzian function only as it is mathemat-ically simpler. For the Gaussian one would proceed in the same way; exceptthat the mathematics dealing with integrals involving the convolution of errorfunctions with harmonic functions and Lorentzians proves tedious. We can,however, infer the quantum results for the Gaussian by the connection betweenthe classical and quantum results derived for the Lorentzian pulse.

We now give the representation of the final state. We choose as ourfinal state of the electromagnetic field one which represents an additionalphoton which is not contained in the phase space spanned by |^tc)« Wedesignate the wave vector of this photon by k^; hence | Yf) can be taken as

1 V Vm> = f 1W ; V (kf>0 ' kl = 0) (C18)

Here | $m> represents the final internal state of the scatterer which wouldcorrespond to m = 0 for Rayleigh processes, and m = 2 for Raman processes.

B. Calculations of the Scattering Intensity

In our calculation we are interested in the one photon emissions; accord-ingly we restrict our consideration to second-order terms in the interactionwhich are treated in the dipole approximation. As such, we use the E • d formof the electromagnetic interaction.

V(t) = -E(t)-d(t) (C19)

115

Page 121: study of resonance light scattering for remote optical probing

The second-order terms to be considered, Eq. (C 6) are of the form

ft ri,

- I / f t 2 / d t a d t x <f jVttaWtOJi) (C20)r* rti' I *, / <*!

-/ _co -' -oo

Inserting the forms for the initial and final states as previously discussed,we have

<f •k X 1 S 10

/•w •*•jkg, Xg

X expr[u)m l t s-(Uojt j+iT l /2( t a- t1)] (C21)

f +a. . a;kf, Xf [ ka.

,„ ox ¥. a. . a; . e S -a, . eic1 k, X ka. X3 ks, X2

[ + itjuk, ti -iu>k til .a, . e Ki -a, . e K* l

k i .X- i ki.^i J '

where

We note that we have heuristically incorporated a lifetime FJ1 for the inter-mediate state. Here I\ is interpreted as the sum of quenching and spontaneousdecay rates. From (C21) we obtain the sum of two terms

•2nft . ,.,— > K-m|v«Vm.

exp i ka| -i|-ul kf

iDl "^1,

"3 1 » "-21 \ Kfexp i ta|«iml

116

Page 122: study of resonance light scattering for remote optical probing

a^ exp i tgjiDnn + j-t! |uu01 - - - (C23)

i t 2 juum i + 0 ) - 1 ! u)01

We are interested in near resonance phenomenon, UUL ~~* ^lo- Hence thedominant contributions from the above terms will be the ones which exhibit anear stationary phase over the full time domains. This corresponds to thelast term in Eq. (C23).

Performing the double-time integral in Eq. (C20), we have

2TT . I1/2. I V^ /^kf l 1 / 3

(C24)

We will proceed with the calculation for the Lorentzian packet case.

One first converts the kz sum to an integral over positive frequency.Then, to order A/UJJ_, relative to unity, the uu integral can be extended to - <=°with the resulting form,

. , -iwt.f - A - W o r n ) / r — / & eJkf,

(C25)

where

K - 9 /- £i I

For t < 0, one closes the contour in the upper-half plane picking up theresidue at the pole uu = uu^, + iA/2,

exp i [cukf - u)om - UJL - i&/ 2] tfk f ,X f <

( t ) [ u ) L - u U i o + i / 2 ( r 1 + A ) ] [uuL-u>kf + u)0m + iA/2] ' (C26)

117

Page 123: study of resonance light scattering for remote optical probing

For t > 0, we close the contour picking up the residues at the three poles inthe lower half of the complex u) plane,

ff kf.

exp - i[u)kf - (

i exp -i[uuL-iA/2]t %

' "(C27)

We desire to compute the average rate of photon production at a time t whichis contained within the frequency interval uuc ± 6/2, within a solid angleelement dn. The frequency interval corresponds to a spectrometer channel.If we represent the probability by dF^f(t)/dn, the production rate is givenby

/

uuc+6/2

"? !<l fk f ,X f( t )l\ve. d*f • (C28)

•u)c-5/2

The averaging bracket ( >ave< in the above equations represents the averageover the initial states of the scatterer. By performing such an average as aseries of convolution integrals over distributions, we can incorporate in-homogeneous broadening effects such as collisional and Doppler broadening.For simplicity in the following we will neglect, for now, the inhomogeneousbroadening.

For t < 0, we obtain

u)c+6/2dF x <( t ) K3V_ = _

dfl (2Tic)U ) - 6 / 2

118

Page 124: study of resonance light scattering for remote optical probing

We note that the absolute magnitudes of the emission and any angular depen-dences of the emission depend upon the evaluation of the matrix elementscontained within K. We will leave this in its general form. We see that thecontribution for t < 0 identically follows the time dependence of the envelopeof the pulse with a coefficient whose magnitude depends upon the proximity toresonance and the position and width of the spectrometer channel. To termsof order 6/uuc « 1 relative to unity, we obtain an emission rate,

6+2(uj -iu_ -Hju_! c L, modtdfi (2TTc)3[(u)-uu7oT

T mn \ I

(C30)

For maximum response, we center the detection channel at u>c = u)L - u>mo

which is the center position of the emitted Raman or Rayleigh lines. For6» A, F, which is the regime of interest, the line is contained within thespectrometer channel and the results obtained for the maximum emission ratewithin the channel are:

K*VA -A|tl' -

I-l

(C3D

Therefore, for the emission rate for t < 0, we are able to derive the resonantenhancement throughout the regime of resonance. For t > 0, the evaluation ofEq. (C28) with the expressions given in Eq. (C27), becomes complicated for thegeneral case. Accordingly, we will investigate two regimes: (a) nearresonance WL ~ 10 > 6» A, T; (b) resonance fluorescence lu^ = u>10. In bothcases we will center the detection channel at the center of the Raman orRayleigh lines, u>c = u)L-uum 0 .

For the case (a) we obtain

- e

dtdfi (2rrc

-At

K-V --L "mo* J ( 2 + e-At ) t a n . l 6 / A (C32)

[n-eA t / 2 ^y sin 6t/2 (1+0(1/6t))] (1+ 0((A/6) a , (Fi/6)3))

For times t » 6"1 ~10~12 sec, our results become

.dtdQ (2n)2c3 (uo10 -uuL)a ' (C33)

119

Page 125: study of resonance light scattering for remote optical probing

Any fluorescent contribution, i. e. , with an exponential decay rate ofexp(-rit) obtained from the evaluation of the integrals is at least of orderA2 / 62, r f / 6 2 relative to unity.

For case (b), UUL = ^xo* resonance fluorescence, the result obtained is

3 3 -4|t| .dtdfi ~(2TT)*c

d*FXf> ( t ) 4K2VA ^L'V/ \ -At /2 2A -TV/21dtdO ~(2n)aca (F i -A) 2 [G " T i + A 6 J

- -G " 6 * (C35)

We see that as we traverse from the near -resonance condition U)L-UO I O >& tothe exact resonance limit uuj_| = u) l 0» the fluorescence part identifiable by adecay of the form exp(-I\t) has increased substantially in its contributionto the emission rate; that is, at near-resonance we start out with an emissionthat follows the packet envelope. As we then move closer into resonanceboth the part which follows the packet envelope and a part which exhibits anexponential decay rate become appreciably larger, with the exponential partgrowing at a faster rate until we achieve the resonance fluorescence conditiongiven in Eq. (C35).

An important point to stress here is that the results obtained in Eqs. (C31)through (C35) for the Lorentzian packet are identical as far as time dependenceand resonant enhancement to that obtained in the classical calculation givenin Chapter II. As we remarked before, the quantum calculation for a Gaussianwave packet contains some difficult integrals; however, now that the connectionbetween the classical and quantum mechanical cases has been determined fora Lorentzian distribution, we can reasonably infer that the same connectionwill prevail for the Gaussian case also.

120

Page 126: study of resonance light scattering for remote optical probing

APPENDIX D

Volume 23, number 1 CHEMICAL PHYSICS LETTERS 1 November 1973

A NEW EFFECT IN THE DEPOLARIZATION OF RESONANT LIGHT SCATTERINGFROM MOLECULES IN THE VAPOR PHASE: I2 *

S.D. SILVERSTEIN and R.L. St. PETERSGeneral Electric Corporate Research and Development, Schenectady, New York 12301. USA

Received 20 July 1973

Theoretical discussions of the depolarization ratios in two types of resonance processes are given. These processesare discrete state resonance fluorescence, and Raman scattering when the resonance is between a bound state andevanescent-dissociative states. Theory predicts a dramatic difference for these two cases due to a quantum mechanicalphase cancellation in the case of virtual dissociative states. The predictions have been verified experimentally.

In this note, we present the results of theoreticaltreatments we have made on the depolarization ratiosfor two types of resonant light scattering processes:resonance fluorescence associated with interelectronicabsorption with a specific vibrational—rotational tran-sition, and "resonant" Raman scattering when the res-onance is between a bound state and an evanescentdissociative state.

Quantum mechanics dictates a dramatic differencefor the depolarization in these two processes. The con-cepts given here are general**; for the sake of experi-mental verification, we specifically apply the theoryto the I2 molecule. The theoretical predictions are apriori in the sense of no adjustable parameters. Fromthe experimental results reported in the literature[2-4], and from further high resolution studies wehave performed [5], we see that theory and experi-ment are in close agreement.

Experiments relevant to the theory have been madeby Holzer et al. [2] and Kiefer and Bernstein [3]. Forincident laser wavelengths above the dissociation limitin I2, they observed depolarization ratios for thefundamental in an overtone sequence of ^0.16 for

* This research has been supported by Contract NAS1-11624 from NASA Langley Research Center.

** The results given here correspond to transitions in symme-tric top molecules for the angular momentum K = 0 inboth states. In ref. [ 1 ] we have further generalized the re-sults to molecules having Q branch absorptions.

the Q branch peak at high resolution and « 0.35 foran area measurement incorporating the 0 and Sbranches. Berjot et al. [4], in their experiments usinga 5017 A* laser excitation on I2 vapor, observed dis-crete state depolarization ratios of* 0.75 at zero for-eign gas pressure, and a reduced depolarization of^ 0.35 at high foreign gas pressure.

The details of the theoretical calculations as orig-inally conceived and performed by one of us (SDS)will be published elsewhere [ 1 ]. Suffice it here to dis-cuss the significant differences between the two res-onance cases, and then to give the numerical resultsfrom the application to I2.

First, the electronic transitions of interest in I2 are2 «^ n0; hence the rotational angular momentummust change by A/ = ± 1 in absorption or emission.This implies that the emission in resonance fluores-cence will be a doublet corresponding to a Q branch(A/ = 0) and either an S (A/ = + 2) or an 0 (A/ = - 2)branch. The depolarization ratios for discrete state res-onance fluorescence are known from Placzek's [6]original calculations. The 0 and S branches give 0.75,while the Q branch value depends upon the J quan-tum number. There are two different Q branch termswhich we denote by Q+ corresponding to the transi-tion sequences /->•/ + 1 -*• J, and Q_ corresponding

* There is a misprint in the caption to Table I of ref. [4]. Thecorrect wavelength is 5017 A, not 5107 A as printed.

140

121

Page 127: study of resonance light scattering for remote optical probing

Volume 23, number 1 CHEMICAL PHYSICS LETTERS 1 November 1973

10 15ROTATIONAL QUANTUM NUMBER J

20 25

Fig. 1. A plot of the two Q branch depolarization ratios as a function of rotational angular momentum quantum number J. TheQ± correspond to cases where intermediate state resonances are with states / ± 1.

to J ->• J - 1 -*• /. The depolarization ratios of Q± de-pend upon the rotational quantum number, and bothasymptotically approach 0.75 in the limit of high /.We note that for the discrete case we are in resonancewith only one of the two terms. The depolarizationratios PQ± go to opposite limits at low/. For example,for /= 0, the perpendicular scattering is zero; hencePQ (J = 0) = 0. On the other hand, for / = 1, PQ(/= 1) = 1. In fig. 1, we plot the depolarization ratiosfor the Q branches as a function of J. We note thatthe results given here are in accord with the low pres-sure experiments [4] at 5017 A, as the dominant lineunder the gain curve of the Ar ion laser is the R(26)62-0 transition which has a high enough / to give adepolarization ratio of ^0.75.

The situation is entirely different for the case whenthe incident laser frequency corresponds to transi-tions to evanescent, dissociative molecular states. The"outgoing" scattering solutions can be expanded in

terms of a complete set of partial waves, and theenergy is independent of the "rotational" quantumnumber J in this expansion. For a given / value, bothQ± contribute and their contributions are summedprior to squaring. The very interesting consequenceof th? quantum mechanics which results in the neweffect described here is that these two terms enterwith opposite sign in the depolarized (perpendicular)component of the emission, hence effecting a cancel-lation and a substantial reduction in the depolariza-tion ratio. As there is a continuum of evanescentstates, one must sum the contributions from all the Jvalues as each has a "continuum resonance".

In an actual experimental situation, the results inthis case are sensitive to the spectrometer exit slitfunction. For example, for the case applicable to verynarrow slits where we get negligible amplitudes of theQ and S branch Raman emissions, even for low 7's,we calculate a Stokes fundamental depolarization

141

122

Page 128: study of resonance light scattering for remote optical probing

Volume 23. number 1 CHEMICAL PHYSICS LETTERS 1 November 1973

0.4

0.3

S 02

O.I

0.0

O.J3

0.125

O.I 10 10 100 1000

SPECTROMETER SLIT WIDTH IN CM"1

Fig. 2. A plot of the predicted, theoretical depolarizationratios for 12 at 400° K as a function of slit width with a tri-angular slit function for resonant scattering from dissociativestates.

ratio for the Q branch at 400°K of = 0.13 for a 1 cm"spectrometer slit*. For I2- being so heavy, the resultsare insensitive to temperature.

In fig. 2, we give the theoretical predictions forthe Stokes fundamental depolarization for the case of"resonance" with the evanescent-dissociative molec-ular states in I2 as a function of slit width for a tri-angular slit function with the slit centered at the J =60 Q branch frequency. The asymptotic value forsmall slit widths is the Q branch only contribution,quoted previously, while as the slit is widened the in-crease corresponds to the added contributions of the0 and S branches. The wide slit asymptote, which isequivalent to an area measurement, predicts a depolar-ization of =« 0.33. The results of our a priori theoryare in very good agreement with experiment [2—5].The calculations include only the contribution ofscattering from molecules in the v = 0 initial state."Hot bands", or scattering from vibrationally excited

* We have experimentally measured a Q branch depolariza-tion ratio of 0.12 ± 0.02 with a 1.6cm"1 slit.

molecules, have the same spectral shape and depolari-zation but are shifted slightly toward shorter wave-lengths [3]. The hot band contribution will not affectthe asymptotes of fig. 2 but may slightly alter the in-termediate region. The relative contributions of thedifferent bands can be determined by numerical cal-culations of the vibrational overlap integrals [7].

The most significant difference between the twotypes of resonance processes described here is thequantum-mechanical cancellation which occurs in thedepolarized component in the scattering from evanes-cent intermediate states. If, however, one had a seriesof pressure broadened discrete states which effectivelymeld into a continuum, the contributions would addincoherently; hence there would be no phase cancel-lation and no anticipated substantial reduction in thedepolarization ratio. This latter conclusion is quiterelevant to the pressure experiments of Berjot et al.[4] and indicates that their high pressure emission islargely due to remnant scattering from excited vibra- 'tional levels via resonance with dissociative states,which is present at all pressures, while the discrete res-onance fluorescent contribution from the groundvibrational level, which is dominant at low pressures,has been quenched out.

References

[1] S.D. Silverstein, Depolarization in the Resonant LightScattering From Molecules in the Vapor Phase withSpecific Application to Ij, to be published.

[2] W. Holzer, W.F. Murphy and H.J. Bernstein, J. Chem.Phys. 52(1970)399.

[3] W. Kiefer and H.J. Bernstein, J. Mol. Spectry. 43 (1972)366.

[4] H. Berjot, M. Jacon and L. Bernard, Can. J. Spectry. 17(1972)60.

[5] R.L. St. Peters and S.D. Silvevstein, General Electric Cor-porate Research and Development Report, to be published.

[6] G. Placzek, Handbuch der Radiologie, 2nd Ed., Vol. 6(H) (1934).

[7] P.P. Williams and D.L. Rousseau, Phys. Rev. Letters 30(1973)951.

142

123

Page 129: study of resonance light scattering for remote optical probing

REFERENCES

1. Melfi, S. H. : Remote Measurements of the Atmosphere Using RamanScattering. Applied Optics, vol.11, no. 1, July 1972, pp. 1605-1610.

2. Grams, G. W. ; and Wyman, C. M. : Compact Laser Radar for RemoteAtmospheric Probing. J. Appl. Meteorology, vol. 11, no. 7, Oct. 1972,pp. 1108-1113.

3. Lapp, M. ; Penney, C. M. ; and Asher, J. A. : Application of Light-Scattering Techniques for Measurement of Density, Temperature, andVelocity in Gasdynamics. ARL-73-0045, Jan. 1973.

4. Holzer, W.; Murphy, W. F. ; and Bernstein, H. J. : Resonance RamanEffect and Resonance Fluorescence in Halogen Gases. J. Chem. Phys.,vol. 52, no. 1, Jan. 1, 1970, pp. 399-407.

5. Penney, C. M.; Goldman, L. M. ; and Lapp, M. : Raman ScatteringCross Sections. Nature, vol. 225, no. 58, Feb. 7, 1972, pp. 110-112.

6. Rudder, Ralph R.; and Bach, David R. : Rayleight Scattering of Ruby-Laser Light by Neutral Gases. J. Opt. Soc. Amer. , vol. 58, no. 9,Sept. 1968, pp. 1260-1266.

7. Berjot, M.; Jacon, M. ; and Bernard, L. : Effet Raman de Resonancepar Excitation dans le Continuum d'Absorption: Sections EfficacesRelatives de Diffusion. Optics Communications, vol. 4, no. 2, Oct.1971, pp. 117-120.

8. Berjot, M.; Jacon, M.; and Bernard, L. : Effet Raman de Resonancepar Excitation dans le Continuum d'Absorption: Importance desTransitions Virtuelles. Optics Communications, vol.4, no. 3,Nov. 1971, pp. 246-248.

9. Jacon, Marcel; Berjot, Maurice; Bernard, Lucien: Optique Moleculaire-Passage Continu de la Fluorescence de Resonance a 1'Effet Raman deResonance. C.R. Acad. Sc. Paris, vol. 273, series B, Nov. 29, 1971,pp. 956-959.

10. Berjot, Maurice; Jacon, Marcel; and Bernard, Lucien: PassageContinu de la Fluorescence de Resonance a I1 Effet Raman de Resonance.Can. J. Spectroscopy, vol. 17, no. 2, March 1972, pp. 60-62.

11. Fouche, D. G.; Herzenberg, A.; and Chang, R. K. : Inelastic PhotonScattering by a Polyatomic Molecule: NO2. J. Appl. Phys., vol. 43,no. 9, Sept. 1972, pp. 3846-3851.

125

Page 130: study of resonance light scattering for remote optical probing

12. Mitchell, Allan Charles Grey; and Zemansky, Mark W. : ResonanceRadiation and Excited Atoms. The Macmillon Co. , 1934.

13. Fouche, D. G. ; and Chang, R. K. : Observation of Resonance RamanScattering below the Dissociation Limit in I2 Vapor. Phys. Rev. Lett,vol. 29, no. 9, Aug. 28, 1972, pp. 536-539.

14. St. Peters, R. L. ; Silverstein, S. D.; Lapp, M; and Penney, C. M. :Resonance Raman Scattering or Resonance Fluorescence in I2 Vapor?Phys. Rev. Lett., vol. 30, no. 6, Feb. 5, 1973, pp. 191-192.

15. Bowman, M. R.; Gibson, A. J. ; and Sandford, M. C.W. : AtmosphericSodium Measured by a Tuned Laser Radar. Nature, vol. 221, Feb. 1,1969, p. 456.

16. Felix, F.; Keenliside, W. ; Kent, G. S.; and Sandford, M.C.W. : LaserRadar Measurements of Atmospheric Potassium. Paper presented atFifth Conference on Laser Radar Studies of the Atmosphere (Williams-burg, Va.) June 4-6, 1973.

17. Glauber, R. J. : Quantum Theory of Coherence, in Quantum Optics,S. S. Kay, and A. Maitland, eds., Academic Press, New York, 1970.

18. Penney, C. M. : Light Scattering in Terms of Oscillator Strengths andRefractive Indices. J. Opt. Soc. Amer. , vol. 59, no. 1, Jan. 1969,pp. 34-42.

19. Griggs, M. : Absorption Coefficients of Ozone in the Ultraviolet andVisible Regions. J. Chem. Phys. vol. 49, no. 2, July 15, 1968,pp. 857-859.

20. Mettee, H. D. : Fluorescence and Phosphorescence of SO2 Vapor.J. Chem. Phys. vol. 49, no. 4, Aug. 15, 1968, pp. 1784-1793.

21. Bethke, George W. : Oscillator Strengths in the Far Ultraviolet. I.Nitric Oxide. J. Chem. Phys., vol. 31, no. 3, Sept. 1959, pp. 662-668.

22. St. Peters, R. L.; and Silverstein, S. D. : Manifestations of PressureBroadening on Tuned Resonance Raman Fluorescence. OpticsCommunications, vol. 7, no. 3, March 1973, pp. 193-196.

23. Herzberg, Gerhard, : Molecular Spectra and Molecular Structure.Vol. II. Infrared and Raman Spectra of Polyatomic Molecules. D.Van Nostrand Co., Inc. (New York), 1945.

126

Page 131: study of resonance light scattering for remote optical probing

24. Hall, Jr. , T.C. ; and Blacet, F. E. : Separation of the AbsorptionSpectra of NOa and N2O4 in the Range of 2400-500 A. J. Chem. Phys.vol. 20, no. 11, Nov. 1952, pp. 1745-1749.

25. Behringer, Josef: Observed Resonance Raman Spectra, pp. 168-223in Raman Spectroscopy. Herman A Szymanski, ed. , Plenum Press,1967.

26. Kiefer, W. ; and Bernstein, H. J. : Vibration-Rotational Structure inthe Resonance Raman Effect of Iodine Vapor. J. Mol. Spectr. , vol.43, no. 3, Sept. 1972, pp. 366-381.

27. Williams, P.F. ; and Rousseau, D. L. : Resonant Raman Scatteringfrom Iodine: Spectral Changes with Excitation Frequency. Phys.Rev. Lett., vol. 30, no. 20, May 14, 1973, pp. 951-954.

28. Placzek, G. : Rayleigh-Streuung und Raman Effekt. ,Handbuch derRadiologie, Vol. 6, Part 2, Akademische Verlagsgesellschaft, 1934.English Trans. UCRL-Trans-526(L).

29. Mrozowski, S. : The Polarization of Band Fluorescence of SomeMetallic Vapors. Acta Phys. Pol. , vol. 7, 1938, pp. 45-48.

30. Pringsheim, Peter: Fluorescence and Phosphorescence. Inter-science Publishers, Inc. (New York), 1963.

31. Silverstein, S.D. ; and St. Peters, R. L. : A New Effect in the Depolari-zation of Resonant Light Scattering from Molecules in the Vapor Phase:I2. Chem. Phys. Letts., vol. 23, no. 1, Nov. 1,1973, pp. 140-142.

32. Stair, Ralph; Schneider, William E.; and Jackson, John K. : A NewStandard of Spectral Irradiance. Applied Optics, vol. 2, no. 11,Nov. 1963, pp. 1151-1154.

33. Anon.: Eastman White Reflectance Paint, Kodak Publication No. JJ-32.

34. Sakurai, Katsumi; and Broida, H. P. : Spectral Study of NC>2 FluoresceneExcited by 11 Lines of Argon and Krypton Ion Lasers. J. Chem. Phys.vol. 50, no. 6, March 15, 1969, pp. 2404-2410.

35. Sackett, Philip B.; and Yardley, James T. : Radiative Lifetime ofNitrogen Dioxide using a Tunable Organic Dye Laser. Chem. Phys.Lett. vol. 6, no. 4, Aug. 15, 1970, pp. 323-325.

127

Page 132: study of resonance light scattering for remote optical probing

36. Sackett, P. B. ; and Yardley, J.T. : Short-Lived Fluorescence fromNitrogen Dioxide. Chem. Phys. Lett. vol. 9, no. 6, June 15, 1971,pp. 612-614.

37. Sackett, Philip B.; and Yardley, James T. : Dynamics of NO2 ElectronicStates Excited by a Tunable Dye Laser. J. Chem. Phys. , vol. 51, no. 1,July 1, 1972, pp. 152-166.

38. Stevens, C. G.; Swagel, M.W. ; Wallace, R. ; and Zare, R. N. : Analysisof Polyatomic Spectra Using Tunable Laser-Induced Fluorescence:Applications to the NC>2 Visible Band System. Chem Phys. Lett. , vol.18, no. 4, Feb. 15, 1973, pp. 465-469.

39. Douglas, A. E. : Anomalously Long Radiative Lifetimes of MolecularExcited States. J. Chem. Phys., vol. 45, no. 3, Aug. 1, 1966, pp.1007-1015.

40. Penney, C. M. ; St. Peters, R. L. ; and Lapp, M. : Absolute Intensity andPolarization of Rotational Raman Scattering from N2, O2, and CO2.NASA CR-121091, 1973.

41. Fouche, D. G.; and Chang, R. K. : Relative Raman Cross Section forO3, CH4, C8H8, NO, N2O and H2. Appl. Phys. Lett. vol. 20, no. 7,April 1, 1972, pp. 256-257.

42. Fouche, D. G. ; and Chang, R. K. : Relative Raman Cross Section ForN2, O2, CO, CO2, SO2, and H2S. Appl. Phys. Letts, vol. 18, no. 12,June 15, 1971, pp. 597-580.

43. Hansch, T. W, : Repetitively Pulsed Tunable Dye Laser for HighResolution Spectroscopy. Appl. Opt. vol. 11, no. 4, April 1972, pp.895-898.

44. Bergman, A. ; David, R.; and Jortner, J. : A Powerful Broad BandTunable Dye Laser. Opt. Comm. , vol. 4, no. 6, Feb./March 1972,pp. 431-433.

45. Singh, S. ; Bonner, W. A.; Potopowicz, J. R.; and van Uitert, L. G. :Nonlinear Susceptibility of Lithium Formate Monohydrate. Appl.Phys. Lett., vol. 17, no. 7, Oct. 1, 1970, pp. 292-294.

46. Washburn, Edward W., ed. : International Critical Tables. Vol. VII.McGraw-Hill, Inc., 1930.

47. Clements, J.H. : On the Absorption Spectrum of Sulphur Dioxide.Phys. Rev., vol. 47, no. 3, Feb. 1, 1935, pp. 224-232.

128

Page 133: study of resonance light scattering for remote optical probing

48. Price, W.C. ; and Simpson, D. M. : Absorption Spectra of SulphurDioxide and Carbon Bisulphide in the Vacuum Ultra-Violet. Proc.Roy. Soc., vol. A165, 1938, pp. 272-278.

49. Metropolis, N. ; and Beutler, H. : Absorption System of Sulphur Dioxideat 3800A. Phys. Rev., vol. 57, June 1, 1940, p. 1078.

50. Metropolis, N. : The Structure of Electronic Bands of PolyatomicMolecules. Phys. Rev., vol. 60, no. 4, Aug. 15, 1941, pp. 283-294.

51. Metropolis, N. : Vibrational Analysis of the Absorption System ofSulphur Dioxide of X 3400-2600. Phys. Rev., vol. 60, no. 4, Aug. 15,1941, pp. 295-301.

52. Merer, A. J. : Rotational Analysis of Bands of the 3800A System of SO2.Discussions of the Faraday Society, no. 35, 1963, pp. 127-136.

53. Warneck, P. ; Marmo, F. F. ; and Sullivan, J.O. : Ultraviolet Absorptionof SO2. Dissociation Energies of SC>2 and SO. J. Chem. Phys. , vol. 40,no. 4, Feb. 15, 1964, pp. 1132-1136.

54. Schotland, R. M.; Chang, D. ; and Bradley, J. : Study of Active Probingof Water Vapor Profiles and Results of Experiments. New YorkUniversity, College of Engineering and Science, Tech. Rep. No. GSL-TR-65-6, 1965.

55. Byer, Robert L. ; and Garbuny, Max: Pollutant Detection by AbsorptionUsing Mie Scattering and Topographic Targets are Retroreflectors.Appl. Opt., vol. 12, no. 7, July 1973, pp. 1496-1505.

56. Kildal, Helge; and Byer, Robert L. : Comparison of Laser Methods forthe Remote Detection of Atmospheric Pollutants. Proc. IEEE, vol.59, no. 12, Dec. 1971, pp. 1644-1663.

57. Dinev, S. G.; Stamenov, K. V.; and Tomov, I. V. : Generation ofTunable UV Radiation in the Range 216-314 nm. Optics Communications,vol. 5, no. 5, Aug. 1972, pp. 419-421.

58. Penner, S. S. : Quantitative Molecular Spectroscopy and Gas Emissivities.Addison-Wesley Publishing Co. , Inc. 1959.

59. Staff of United States Air Force Air Research and Development Command,Air Force Research Division, Geophysics Research Directorate:Handbook of Geophysics. Macmillian (New York), 1960.

129

Page 134: study of resonance light scattering for remote optical probing

60. Mavrodineanu, Radu; and Boiteux, Henri: Flame Spectrscopy. JohnWiley and Sons, Inc., 1965.

61. Hibben, J. H. : The Raman Effect and its Chemical Applications.Reinhold Press (New York), 1939.

62. Goldberger, M.~L.; and Watson, K. M.: Collision Theory. OxfordUniversity Press, 1954.

63. For a recent review article with additional references, see: Glauber,R. J. : Quantum Theory of Coherence, in Quantum Optics. Kay, S.H.,and Maitland, A., editors, Academic Press, New York, 1970.

130

Page 135: study of resonance light scattering for remote optical probing

NASA CR-132363

DISTRIBUTION LISTNAS1-11624

No.Copies

NASA Langley Research CenterHampton, VA 23665Attn: Report & Manuscript Control Office, Mail Stop 180A 1

Raymond L. Zavasky, Mail Stop 115 1Wendell G. Ayers, Mail Stop 217K 1Dr. Frank Allario, Mail Stop 193A 1Dr. Reginald J. Exton, Mail Stop 2 35A 1Dr. G. Burton Northam, Mail Stop 401A 10

NASA Ames Research CenterMoffett Field, CA 94035Attn: Library, Mail Stop 202-3 1

NASA Flight Research CenterP.O. Box 273Edwards, CA 93523Attn: Library 1

NASA Goddard Space Flight CenterGreenbelt, MD 20771Attn: Library 1

Richard J. Pinamonti, Jr. 1

NASA Lyndon B. Johnson Space Center2101 Webster Seabrook RoadHouston, TX 77058Attn: Library, Code JM6 1

M. Jay Harnage, Jr., Code HC 1Dr. Robert D. Hudson, Code TN2 1

Jet Propulsion Laboratory4800 Oak Grove DrivePasadena, CA 91103Attn: Library, Mail 111-113 1

NASA Wallops StationWallops Island, VA 23337Attn: Library 1

Alfred C. Holland, Code DAS, Building E-106 1

Page 136: study of resonance light scattering for remote optical probing

D2

Copies

NASA Lewis Research Center21000 Brookpark RoadCleveland, OH 44135Attn: Library, Mail Stop 60-3 1

Tom A. Coney, Mail Stop 500-318 1Jack A. Salzman, Mail Stop 500-318 1

NASA Marshall Space Flight CenterHuntsville, AL 35812Attn: Library 1

Kenneth A. Kadrmas, S&E-AERO-F 1

NASA John F. Kennedy Space CenterKennedy Space Center, FL 32899Attn: Library, IS-DOC-1L 1

National Aeronautics & Space AdministrationWashington, DC 20546Attn: KSS-10/Library 1

ERF/Michael A . Calabrese , 1ERF/Jules Lehmann 1ERF/Luke L. Liccini 1REE/Dr. Bernard Rubin 1

Environmental Protection AgencyNational Environmental Research CenterP.O. Box 15027Las Vegas, NV 89114Attn: Dr. S. Harvey Melfi 1

Environmental Protection AgencyResearch Triangle Park, NC 27711Attn: Bill Herget 1

United Aircraft CorporationUnited Aircraft Research LaboratoriesEast Hartford, CT 06108Attn: Bill Morey 1

Stanford University,Department of PhysicsMenlo Park, CA 94305Attn: Dr. Bob Byers 1

Page 137: study of resonance light scattering for remote optical probing

D3

Copies

Block Engineering, Inc.19 Blackstone StreetCambridge, MA 02139Attn: Stanley Klainer 1

Dr. Rudolf Penndorf148 Oakland StreetWellesley Hills, MA 02181 1

New York UniversityUniversity HeightsBronx, NY 10453Attn: Professor Richard M. Schotland 1

University of MissouriRolla, MO 65581Attn: Dr. Burton Schuster 1

Mitre CorporationMcLean, VA 22067Attn: Edward A. Ward 1

University of Wisconsin1225 Dayton StreetMadison, WI 53706Attn: Dr. Edwin Eloranta 1

James Weinman 1

Stanford Research Institute333 Ravens wood AvenueMenlo Park, CA 94025Attn: Edward K. Proctor, Jr. 1

R. T.H. Collis 1

Air Force Cambridge Research LaboratoriesBedford, MA 01731Attn: Dr. Louis Elterman 1

Dr. Frederic E. Volz (OPA) 1

McDonnell Douglas CorporationMcDonnell Douglas Aerospace Center5301 Bolsa AvenueHuntington Beach, CA 92647Attn: Dr. Jack Grossman, A3-833 1

Page 138: study of resonance light scattering for remote optical probing

D4

Copies

Drexel UniversityPhiladelphia, PA 19104Attn: Professor John A. Cooney 1

Massachusetts Institute of TechnologyLincoln LaboratoryP. O. Box 73Lexington, MA 02173Attn: Dr. Daniel G. Fouche . 1

D r . Paul L . Kelly . 1

University of ArizonaTucson, AZ 85721Attn: Dr. Ben Herman 1

Edge wood Arsenal, MD 21010Attn: Harvey Tannenbaum 1

College of the City of New YorkConvent Avenue and W 139 StreetNew York, NY 10031Attn: Professor Samir A. Ahmed 1

The Aerospace CorporationElectronics Research LaboratoriesWuantum Optics DepartmentP.O. Box 92957'Los Angeles, CA 90045Attn: Dr. Milton Birribaum 1

Martin Marietta CorporationThe Martin CompanyDenver DivisionPost Office Box 179Denver, CO 80201Attn: Nicholas J. Ganiaris 1

National Center for Atmospheric ResearchBoulder, CO 80302Attn: Dr. Gerald W. Grams 1

Page 139: study of resonance light scattering for remote optical probing

D5

Copies

Tohoku UniversityResearch Institute of Electrical CommunicationKatahira 2-ChomeSendai 980, JapanAttn: Professor Humio Inaba 1

Canada Center for Remote SensingDepartment of Eenergy Mines & ResourcesOttawa, Ontario klAOE4CanadaAttn: Dr. Michael Bristow ; 1

NASA Scientific & Technical Information FacilityP. O. Box 33College Park, MD 20740 12

plus reproducible