We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF or read online on Scribd
You are on page 1/ 95
274
FEMTOCHEMISTRY
ATOMIC-SCALE DYNAMICS OF THE CHEMICAL BOND
USING ULTRAFAST LASERS
Nobel Lecture, December 8, 1999
by
Aumep H. Zewat
California Institute of Technology, Pasadena, USA.
Over many millennia, humankind has thought to explore phenomena on an
ever shorter time scale. In this race against time, femtosecond resolution (1 fs
= 101 second) is the ultimate achievement for studies of the fundamental dy-
namics of the chemical bond. Observation of the very act that brings about che-
mistry - the making and breaking of bonds on their actual time and length
scales — is the wellspring of the field of Femlochemistry, which is the study of
molecular motions in the hitherto unobserved ephemeral transition states of
physical, chemical and biological changes. For molecular dynamics, achieving
this atomic-scale resolution using ultrafast lasers as strobes is a triumph, just as
x-ray and electron diffraction, and, more recently, STM and NMR, provided
that resolution for static molecular structures. On the femtosecond time
scale, matter wave packets (particle-type) can be created and their coherent
evolution as a single-molecule trajectory can be observed. The field began with
simple systems of a few atoms and has reached the realm of the very complex
in isolated, mesoscopic and condensed phases, and in biological systems such
as proteins and DNA. It also offers new possibilities for the control of reac
tivity and for structural femtochemistry and femtobiology. This anthology
gives an overview of the development of the field from a personal perspective,
encompassing our research at Caltech and focusing on the evolution of tech-
niques, concepts, and new discoveries.Ahmed H. Zewail 275
TABLE OF CONTENTS:
Abstract
L
IL.
Mm.
Prologue
Dynamics and Arrow of Time
(1) Origins - From Kinetics to Dynamics
‘The Arrhenius Seminal Contribution
The London, Eyring and Polanyi Contributions
‘Transition State and Definition
‘Transition State and Spectroscopy
(2) Arrow of Time
Femtochemistry: Development of the Field
(1) The Early Years of Coherence
New Techniques for Molecules
Optical Analogue of NMR: Controlling the Phase
(2) The Marriage with Molecular Beams
The Anthracene Discovery: A Paradigm Shift
‘The Successful 036 Laboratory
Changing A Dogma: Development of RCS
ransition to the Sub-Picosecond Regime
‘A New Beam Machine: Pump-Probe Mass Spectrometry
The First ICN Experiment: Sub-picosecond Resolution
(4) The Femtosecond Dream
A Piece of Good Fortune
‘The Classic Femtosecond ICN Discovery
‘The Nal Discovery: A Paradigm for the Field
The Saddle-Point Transition State
‘The Uncertainty Principle Paradox
Bimolecular, Bond Making & Breaking: Bernstein’s Passion
Ultrafast Electron Diffraction (UED)
Clusters, Dense Fluids & Liquids, and New Femtolands
Theoretical Femtochemistry
Experimental Femtochemistry
(5) Femtocopia ~ Examples from Caltech
Elementary Reactions & Transition States
Organic Chemistry
Electron & Proton Transfer
Inorganic & Atmospheric Chemistry
The Mesoscopic Phase: Clusters & Nanostructures
The Condensed Phase: Dense Fluids, Liquids & Polymers276 Chemistry 1999
(6) Opportunities for the Future
Transient Structures from Ultrafast Electron Diffraction
Reaction Control
Biological Dynamics
IV. Impact and Concepts - A Retrospective
‘Time Resolution — Reaching the Transition-State Limit
Atomic-Scale Resolution
Generality of the Approach
Some Concepts
(1) Resonance - Non-equilibrium Dynamics
(2) Coherence - Single-molecule-type Dynamics
(3) Transition Structures - Landscape Dynamics
(4) Reduced Space — Directed Dynamics
V. Epilogue
Acknowledgments
References
Bibliography:
The Caltech Research
Some General References
I. PROLOGUE
In the history of human civilization, the measurement of time and recording
the order and duration of events in the natural world are among the earliest
endeavors that might be classified as science. The development of calendars,
which permitted the tracking of yearly flooding of the Nile valley in ancient
Egypt and of the seasons for planting and harvesting in Mesopotamia, can be
traced to the dawn of written language. Ever since, time has been an impor-
tant concept [1] and is now recognized as one of the two fundamental di-
mensions in science. The concept of time encapsulates an awareness of its
duration and of the passage from past to present to futureand surely must have
existed from the very beginning with humans searching for the meaning of
birth, life and death and, in some cultures, rebirth or recurrence.
My ancestors contributed to the beginning of the science of time, develop-
ing what Neugebauer [1] has described as “the only intelligent calendar which ever
existed in human history”. The “Nile Calendar” was an essential part of life as it
defined the state of yearly flooding with three seasons, the Inundation or
Plooding, Planting, and Harvesting, each four months long. A civil year lasting
365 days was ascertained by about 3000 BC or before, based on the average
time between arrivals of the flood at Heliopolis, just north of Cairo; nilome-
ters were used in more recent times, and some are still in existence today. By
the time of the First Dynasty of United Egypt under Menes in ca. 3100 BG, the
scientists of the land introduced the concept of the “Astronomical Calendar”
by observing the event of the helical rising of the brilliant star Sothis (orAhmed H. Zewail 277
Sirius). Inscribed on the Ivory Tablet, dating from the First Dynasty and now
at the University Museum in Philadelphia, were the words, “Sothis Bringer of
the Year and of the Inundation” [1, Winlock]. On the Palermo Stone, the an-
nals of the kings and their time-line of each year’s chief events were document-
ed from pre-dynastic times to the middle of the Fifth Dynasty [1, Breasted].
Thus, as early as 3100 BG, they recognized a definite natural phenomenon
for accurate timing of the coming flood and recounted the observed reap-
pearance of the star as the New Year Day — real-time observation of daily and
yearly events with the zero-of time being well defined!’
In about 1500 BG, another major contribution to this science was made,
the development of Sun-Clocks, or Sundials, using moving shadows (Fig. 1].
Now in Berlin, the sun-clock bearing the name of Thutmose III (or Thothmes
after the Egyptian God Thoth of wisdom and enlightenment), who ruled at
Thebes from 1447-1501 BC, shows the graduation of hours for daytime
measurements. This clock with uneven periods for hours was man-made and
transportable. For night time, the water-clock was invented, and the device
provided even periods for timing. With these developments, the resolution of
time into periods of year, month, day and hour became known and has now
been used for more than three millennia. The durations of minutes and se-
conds followed, using the Hellenistic sexagesimal system, and this division, ac-
cording to Neugebauer, “is the result of a Hellenistic modification of an Egyptian
practice combined with Babylonian numerical procedures.” About 1300 AD, the me-
chanical clock [1, Whitrow] was advanced in Europe, ushering in a revolution
in precision and miniaturization; Galileo began studies of pendulum motions
and their clocking, using his heartbeats (seconds), in 1582 AD. Our present
time standard is the cesium atomic clock’, which provides precision of about
1:10", ie., the clock loses or gains one second every 1.6 million years. For this
work, Norman Ramsey shared the 1989 Nobel Prize in Physics.
Until 1800 AD, the ability to record the timing of individual steps in any
process was essentially limited to time scales amenable to direct sensory per
ception - for example, the eye’s ability to see the movement of a clock or the
Recognizing the incommensurability of the lunar month and the solar year, they abandoned
the lunar month altogether and used a 30-day month, Thus the year is made of 12 months, 30
days each, with 5 feast days at the end of the year. The “Civil Calendar” was therefore 365 days per
year and differs from the astronomical calendar of Sothis by approximately a 1/4 day every year.
The two calendars must coincide at intervals of 365 x 4=1460 years, and historians, based on re-
corded dates of the reappearance of Sothis in dynastic periods, give the four dates for the coin-
cidence of both calendars: 139 AD, 1317 BC, 2773 BC, and 4233 BC (+139-3 x 1460 = 4241 BC).
Even though the Egyptians discovered the astronomical calendar, 365-1/4 days, they decided,
presumably for “bookkeeping”, to use the civil calendar of 365 without leap years. They also
divided the day into two periods of 12 hours each for day and night time. The remarkable calen-
dar of years, months, days, and hours was adapted throughout history and formed the basis for
the 365.25-day Julian (46 BC, Alexandria school) and 365.2422-day Gregorian (from 1582 AD,
Pope Gregory XIII) calendars. In the words of the notable Egyptologist J. H. Breasted [1], “It has
thus been in use uninterruptedly over six thousand years,” Many historians regard the date around
4241 BC as the beginning of history itself as it defines the period of written records; anything
earlier is prehistory
2 Since 1967 one second has been defined as the time during which the cesium atom makes ex-
actly 9, 192,631,770 oscillations.278 Chemistry 1999
Glendar Sundial Mechanical Cock Fast Photography, Laver Femioscopy:
teal me/ard, — (Ponptawant (icone bla sa ai
Figure 1, Timeline of some events in the history of measurements of time, from yearly calendars
to the femtosecond regime (see text).
car’s ability to recognize a tone. Anything more fleeting than the blink of an
eye (~0.1 second) or the response of the ear (~0.1 millisecond) was simply be-
yond the realm of inquiry. In the nineteenth century, the technology was to
change drastically, resolving time intervals into the sub-second domain, The
famous motion pictures by Eadweard Muybridge (1878) of a galloping horse,
by EtienneJules Marey (1894) of a righting cat, and by Harold Edgerton
(mid-1900’s) of a bullet passing through an apple and other objects are ex-
amples of these developments, with millisecond to microsecond time resolu-
tion, using snapshot photography, chronophotography and stroboscopy, re-
spectively [2]. By the 1980's, this resolution became ten orders of magnitude
better [see Section 112], reaching the femtosecond scale’, the scale for atoms
and molecules in motion.
The actual atomic motions involved in chemical reactions had never been
observed in real time despite the rich history of chemistry over two millennia
[3], as khem became khemia, then alchemy, and eventually chemistry [3,4] —
see the Stockholm Papyrus in Fig. 2. Chemical bonds break, form, or geome-
trically change with awesome rapidity. Whether in isolation or in any other
phase, this ultrafast transformation is a dynamic process involving the me-
chanical motion of electrons and atomic nuclei. The speed of atomic motion
°"The prefix milli comes from Latin (and French), micro and nano from Greek, and pico from
Spanish, Femto is Scandinavian, the root of the word for “fifteen”; nuclear physicists call femto-
meter, the unit for the dimensions of atomic nuclei, fermi. Ato is also Scandinavian.Ahmed H. Zewail 279
Figure 2. A page of the Stockholm papyrus (300 AD) describing the recipe for “making” (actual-
ly imitating) emerald. Note that change has been at the heart of chemistry from its millennia-old
definition KM, to Chémia and Alchemy, and to Chemistry. [Ref. B26, and references therein]
Equilibrium, kinetics and dynamics are the foundation for the describtion of chemical changes.
is ~ 1 km/second and, hence, to record atomic-scale dynamics over a distan-
ce ofan Angstrm, the average time required is ~ 100 fs. The very act of such
atomic motions as reactions unfold and pass through their transition states is
the focus of the field of femtochemistry. With fs time resolution we can “free-
ze” structures far from equilibrium and prior to their vibrational and rotatio-
nal motions, or reactivity. The pertinent questions about the dynamics of the
chemical bond are the following: How does the energy put into a reactant
molecule redistribute among the different degrees of freedom, and how fast
does this happen?; What are the speeds of the chemical changes connecting
individual quantum states in the reactants and products?; What are the detai-280 Chemistry 1999
semicon |
| Plosiew). Cheintesh and Blalegtaal | soso btocas Mode |
Figure 3, Time scales. The relevance to physical, chemical, and biological changes. The funda
mental limit of the vibrational motion defines the regime for femtochemistry. Examples are
given for each change and scale. [Ref. B4, B10]
led nuclear motions that chart the reaction through its transition states, and
how rapid are these motions? As pointed out by Jim Baggot, “the entire history
of chemical reaction dynamics and kinetics has been about providing some approxima-
te answers to these three questions [4].”
In femtochemistry, studies of physical, chemical or biological changes are
the fundamental time le of molecular vi
tions [Fig. 3]. In this sense, femtoscience represents the end of the race
against time, or, as the report of Ref. [5] puts it, ”.
For this same reason, Martens stated, “all chemistry is femtochemistry [5]". The
ephemeral transition states, denoted in the past by a bracket [TS]* for th
clusiveness, can now be clocked as a molecular species TS!. Moreover, on this
time scale, the time-dependent description of a coherent, single-molecule tra-
jectory represents the classical nuclear “motion picture” of the reaction as its
wave packet proceeds from the . and on
to final products ~ the language of the actual dynamics! The fs time scale is
unique for the creation of such coherent matter waves on. the atomic scale of
length, a basic problem rooted in the development of quantum mechanics
and the duality of matter. Figure (4) highlights some steps made in the de-
scription of the duality of lightmatter and time scales, and the Appendix
cusses the importance of coherence for localization,
the actual nuclear
tions
reaching the end of the road”.
nitial state, through transition statesAhmed H. Zewail 281
Matter Waves
Particle-type Control & Dynamics
de Broglie (1924)
Binstein’s light wave/particle
E=hv E=cp
i=hlp
Similarly, matter particlevwave
Schrodinger (1926)
‘The Wave Equation ~ Stationary waves
HY =EY
Schrodinger (1926)
Micro- to Macro-mechanies
‘Quantum to Newton Mechanics
¥ to wave group
Femtochemistry & Quantum Limit (i):
Partiele-type
ge Bogie (initial localization) = h/p
uncertainty in time measurement
AxAp2WGn) ALAE>h/Qn)
for force free
Ax=pim Ai
At
At~10fs Ax~O1A
At~10 ps Ax~100A
Figure 4. Matter waves and particle-type limit of dynamics. The atomicscale de Broglie wave-
length in the coherent preparation of a quantum system and in the uncertainty of probe meas
rements are both reached on the fe time scale. Sec also the Appendix and its Figure.
Figure 5. A femtochemistry apparatus typical of early Femtolands. (A, above) The laser system.
(top) the first CPM oscillator used in Femtoland I; (bottom) the continuum generation to the
right and the experimensal layout for clocking to the left. (B, next page) The molecular beam
apparatus of Femtoland IIT, together with a view of the beam/laser arrangement, (Ref. B1, B16,
B28, 49]282 Chemistry 1999
Molecular
beam
Femtosecond
Mice Probing Pulse
This powerful concept of coherence lies at the core of femtochemistry and
was a key advance in observing the dynamics at atomic resolution. The realiz-
ation of its importance and its detection by selectivity in both preparation and
probing were essential in all studies, initially of states and orientations, and
culminating in atomic motions in reactions. With these concepts in mind, the
marriage of ultrafast lasers with molecular beams [Fig. 5A, B] proved to be es-
sential for the initial development. Laser-induced fluorescence was the first
probe used, but later we invoked mass spectrometry and nonlinear optical
techniques. Now numerous methods of probing are known [Fig. 6] and used
in laboratories around the world; Coulomb explosion is the most recent po-
werful probe developed by Will Castleman [5] for arresting reactive interme-
diates, as mentioned in Section IIL
Applications of femtochemistry have spanned the different types of chemi-
cal bonds ~ covalent, ionic, dative and metallic, and the weaker ones, hydro-
gen and van der Waals bonds. The studies have continued to address the vary-
ing complexity of molecular systems, from diatomics to proteins and DNA.Ahmed H, Zewail 283
Laser-induced Fluorescence
Infrared Pumping
Laser Multiphoton Ionization
Femtochemistry
‘Mass Spectrometry
Stimulated Emission Pumping) | Frequency Ms
‘Time, Velocity & Angular Resolutions
‘Transient Absorption
Velocity | Orientation
Maltiple Pulses a Photoelectron Spectroscopy
I (Came, ZEKE, Coincidence)
Coulomb Explosion
Diffraction ?
Figure 6, Techniques for probing in femtochemistry. Both excited and ground states have been
probed by these methods. The correlations of time with frequency, mass, velocity and orientation
were essential in the studies of complex systems. Diffraction represents the new effort for prob-
ing structures (see text and Fig. 29).
Studies have also been made in all phases of matter: gases and molecular be-
ams; mesoscopic phases of clusters, nanostructures, particles and droplets;
condensed phases of dense fluids, liquids, solids, surfaces and interfaces; and
in sibling fields of femtoscience such as femtobiology [Fig. 7].
Twenty-four centuries ago, the Greek philosopher Democritus and his
teacher Leucippus gave birth to a new way of thinking about matter’s invisible
and elementary entity, the atom. Richard Feynman once asked, if you had on-
ly one sentence to describe the most important scientific knowledge we pos-
sess, what would that sentence be? He said, “everything is made of atoms.”
Democritus’ atomism, which was rejected by Aristotle, was born on a purely
philosophical basis, surely without anticipating some of the twentieth centu-
ry's most triumphant scientific discoveries. Atoms can now be seen, observed
in motion, and manipulated [6]. These discoveries have brought the micro-
scopic world and its language into a new age, and they cover domains of
length, time and number. The length (spatial) resolution, down to the scale of
atomic distance (angstrém), and the time resolution, down to the scale of
atomic motion (femtosecond), have been achieved. The trapping and spec-
troscopy of a single ion (electron) and the trapping and cooling of neutral
atoms have also been achieved. All of these achievements have been recogniz-
ed by the awarding of the Nobel Prize to STM (1986), to single-electron and
-ion trapping and spectroscopy (1989), to laser trapping and cooling (1997),
and to laser femtochemistry (1999).284 Chemistry 1999
Femtochemistry
‘Structures
UED & X-rays
Femtobiology
Elementary Phenomena: Transition States, Energy Redistributions (IVR) & Rates (k(E))
Chemical Bond: Covalent, Ionic, Dative, Metallic, Hydrogen & van der Waals,
Molecular Size: ‘Two Atoms to DNA & Proteins
Molecular State: Ground & Excited, Neutral & lon (negative or positive)
Figure 7. Femtochemistry and the scope of applications.
IL DYNAMICS AND ARROW OF TIME
(1) Origins ~ From Kinetics to Dynamics
The Arrhenius Seminal Contribution
At the turn of the 20th century, the study of reactivity was dominated by the
question’: How do reactions proceed and what are their kinetic rates? Svante
Arrhenius [7] gave the seminal description of the change in rates of chemical
“The main focus of prior studies was on the thermodynamics and equilibrium characteristics.
Interestingly, chemical equilibrium was already established a century before (1798) by Claude
Berthollet, during a visit to the Natron Lakes in Napoleonic times, through his studies of the
reaction Na,CO, + CaCl, + 2NaCl + CaCO, [S. W. Weller, Bull. Hist, Chem. 24, 61 (1999)].Ahmed H. Zewail 285
reactions with temperature and formulated in 1889 the familiar expression
for the rate constant,
k=A exp (E,/RT) a)
which, as Arrhenius acknowledged, had its roots in van’t Hoff’s (1884) equa-
tions [7]. For any two reactants, the rate constant (k) depends on tempera-
ture (T) according to an energy of activation (E,), which is different from the
thermodynamic net energy change between reactants and products, and the
dependence is exponential in form. (R is the universal gas constant). If we
think of the reaction as a finite probability of reactant A colliding with B, then
the rate is simply the collision frequency times the fraction of successful colli-
sions with an energy equal to or more than E,. Besides the value of equation
(1) through the well-known plots of “In k vs. 1/T” to obtain the energy of ac-
tivation E,, Arrhenius introduced a “hypothetical body’, now known as the
“activated complex”, a central concept in the theory of reaction rates: the
reaction, because of collisions or other means, proceeds only if the energy is,
sufficient to exceed a barrier whose energy is defined by the nature of the
complex. Since then, various experimental data for different temperatures T
were treated with equation (1), yielding E, and the pre-exponential factor A.
A few years after Arrhenius’ contribution, Bodenstein (1894) (7] published
a landmark paper on the hydrogen/iodine system, which has played an im-
portant role in the development of gas-phase chemical kinetics, with the aim
of understanding elementary reaction mechanisms. In the twenties, Linde-
mann (1922), Hinshelwood (1926), Tolman (1920), and others [7&8] devel-
oped, for unimolecular gas-phase reactions, elementary mechanisms with dif
ferent steps describing activation, energy redistribution, and chemical rates.
By 1928, the Rice-Ramsperger-Kassel (RRK) theory was formulated, and
Marcus, starting in 1952, blended RRK and transition state theory in a direc-
tion which brought into focus the nature of the initial and transition-state vi-
brations in what is now known as the RRKM theory [8].
The rate constant, k(T), does not provide a detailed molecular picture of
the reaction. This is because k(T), which was obtained from an analogy with
van't Hoff’s description of the change with T of the equilibrium constant K
(thermodynamics), is an average of the microscopic, reagent-state to product-
state rate coefficients over all possible encounters. These might include dif-
ferent relative velocities, mutual orientations, vibrational and rotational
phases, and impact parameters. A new way was needed to describe, by some
quantitative measure, the process of the chemical reaction itself: How reagent
molecules approach, collide, exchange energy, sometimes break bonds and
make new ones, and finally separate into products. Such a description is the
goal of molecular reaction dynamics [9]-
The London, Eyring and Polanyi Contributions
For some time, theory was ahead of experiment in studies of microscopic mo-
lecular reaction dynamics. The effort started shortly after the publication of
the Heitler-London quantum-mechanical treatment (1927) of the hydrogen286 Chemistry 1999
molecule [10], a breakthrough in thinking not only about the stable structure
of the chemical bond, but also about how two atoms can interact at different
separations. One year later (1928), for Sommerfeld’s Festschrift (60th birth-
day), London [10] presented an approximate expression for the potential
energy of triatomic systems, e.g., H, in terms of the coulombic and exchange
energies of the “diatomic” pairs, In’ 1931 Henry Eyring and Michael Polanyi
[10], using the London equation, provided a semiempirical calculation of a
potential energy surface (PES) of the H + Hy reaction describing the journey
of nuclei from the reactant state of the system to the product state, passing
through the crucial transition state of activated complexes. The birth of
“reaction dynamics” resulted from this pioneering effort and, for the first
time, one could think of the PES and the trajectories of dynamics on it ~ in
those days, often, expressed in atomic units of time! But no one could have dreamed
in the 1930's of observing the transient molecular structures of a chemical
reaction, since the time scale for those far from equilibrium activated complexes
in the transition state was estimated to be less than a picosecond (ps).
The time scale was rooted in the theory developed for the description of re-
action rates. Building on Arrhenius’ work and the work of Polanyi and
Wigner (1928) [11], in 1935, Eyring, and independently Evans and Polanyi,
formulated transition-state theory, which gave an explicit expression for
Arrhenius’ pre-exponential factor [11]:
a
k KT gs KD. Q
h h 0,05
where k is Boltzmann's constant, h is Planck’s constant and Q is the partition
function. The key idea here is to assume the equilibration of the population
between reactants and the transition state: A + B «> [TS]? > products. Thus,
the rate constant can be related to the equilibrium constant for formation of
the transition state, K*, and hence AG?, AH? and ASt; physically, AGt
(AH*-TAS*) in the exponent gives the barrier energy (through AH?) along
the reaction coordinate, and the pre-exponential entropic term which re-
flects the change in vibrational modes perpendicular to the reaction coordi-
nate. By comparing (2) with Arrhenius’ equation (1), A can be identified; E,,
the activation energy, and E,, the barrier energy, with zero-point energy cor-
rections, are related [8, Laidler]. Kramers’ (1940) classic work [11] modified
the pre-exponential factor to include friction from the surrounding medium,
with transition state theory giving an upper limit, and Casey Hynes, in the
1980's, provided a dynamical theory of friction with emphasis on time scales
in the transition-state region and for solvent interaction [11]. In the 1950’s,
Marcus [11] obtained a transition-state type expression for reactions of elec-
tron transfer in solutions with the Gibbs free energy of activation expressing
the dependence on solvent reorganization energies. This work was awarded
the 1992 Nobel Prize.
According to transition state theory, the fastest reaction is given by KT/h,
which is basically the “frequency” for the passage through the transition state
exp(—Ey/kT) (2)Ahmed H. Zewail 287
[Equation 2]. At room temperature this value is 6 x 102 second", correspond-
ing to ~ 170 fs; the time scale of molecular vibrations is typically 10-100 fs.
This estimate is consistent with knowledge of the speeds of nuclei and the
distance change involved in the reaction. In 1936 the first classical trajectory
from Hirschfelder-Eyring-Topley molecular dynamics simulations of the
H+H, reaction showed the fS steps needed to follow the reaction profile,
albeit on the wrong PES. Later, Karplus, Bunker and others showed a range
for the time scales, ps to fs, depending on the reaction and using more
realistic PES’s [see 9 and 26].
Transition State and Definition
In general, for an elementary reaction of the type,
A+BC> [ABC]* > AB+C (3)
the whole journey from reagents to products involves changes in internuclear
separation totaling ~10A. If the atoms moved at 10*-10° cm/sec then the en-
tire 10A trip would take 1012-10 sec. If the ‘transition state’, [ABCI}, is de-
fined to encompass all configurations of ABC significantly perturbed from the po-
lentialenergy of the reagents A + BC or the products AB + C, then this period of
1-10 ps is the time available for its observation. To achieve a resolution of
~0.1A, the probe time window must be 10-100 fs.
The above definition of the transition state follows the general description
given by John Polanyi and the author [12], namely the full family of configur-
ations through which the reacting particles evolve en route from reagents to
products, This description may seem broad to those accustomed to seeing the
TS symbol, t, displayed at the crest of the energy barrier to a reaction. As
stated in Ref. 12, even if one restricts one’s interest to the over-all rates of che-
mical reactions, one requires a knowledge of the family of intermediates
sampled by reagent collisions of different collision energy, angle and impact
parameters. The variational theory of reaction rates further extends the
range of TS of interest, quantum considerations extend the range yet further,
and the concern with rates to yield products in specified quantum states and
angles extends the requirements most of all. A definition of the TS that em-
braces the entire process of bond-breaking and bond-making is therefore
likely to prove the most enduring. This is specially important as we address
the energy landscape of the reaction, as discussed in Section IV.
The cardinal choice of the transition state at the saddle point, of course,
has its origin in chemical kinetics — calculation of rates - but it should be re-
membered that this is a mathematical ‘single-point’ with the division made to
define the speed of a reaction [8]. Even in thermal reactions, there is enough
of an energy distribution to ensure many types of trajectories, Furthermore,
transition state theory is not a quantum theory, but a classical one, because of
the assertion of a deterministic point. Quantum uncertainty demands some de-
localization [8, Hase] and, as mentioned above, the theory invokes the equi-
libration of reactants and transition state populations, using statistical ther-
modynamics. In fact, the term transition state is used ambiguously to refer288 Chemistry 1999
both to the quasi-equilibrium state of the reaction and to the molecular struc.
ture of the saddle point. As discussed in Ref. 13 [Williams, Doering, Baldwin],
the molecular species at the saddle point perhaps could be referred to as the
transition structure; the activated complex is even more descriptive.
The location at the saddle point provides the highest energy that must
be reached, defining the exponential probability factor; the dynamics (forces
and time scales) are governed by the nature of the TS region. Provided
that the energy landscape of the reaction is controlled by a narrow region,
the structure of the transition state becomes important in structure-reactivity
correlations; rates vs. Gibbs energy between the TS and the ground state. (It
is also useful in designing TS analogs as enzyme inhibitors and TS comple-
ments as catalysts [13]). The position of the TS and its energy relative to that
of reactants and products along the reaction path becomes relevant. In the
mid 1930's, the Bell-Evans-Polanyi principle gave a predictive correlation
between changes in barrier heights and the enthalpies of reactions,
especially for a series of related reactions; the TS for an exoergic or an en-
doergic reaction is very different. In 1955, this led to Hammond’s postulate
which characterizes a reactantlike TS (so called “early” TS) for exoergic reac-
tions, productlike TS (so called “late” TS) for endoergic reactions, and “cen-
tral” TS for energy neutral reactions [8, Shaik]. John Polanyi, the son of
Michael, formulated some concepts regarding energy disposal in relation to
the position of the TS on the PES [14], using molecular dynamics simulations
and experimental studies of chemiluminescence.
It should be recognized that selectivities, efficiencies, and stereochemistries
are quantified only when the dynamics on the global energy landscape are
understood. For example, the picture for simple reactions of a few (strong)
bonds being made and broken is changed when the energy surface is nearly
flat or there is a significant entropic contribution. Many transition states will
exist in the region as in the case of protein folding (13, Fersht]. Finally, in the
transition state, chemical bonds are in the process of being grade and broken.
In contrast, for intermediates, whose bonds are fully formed, they are in po-
tential wells, typically “troughs” in the TS region. However, the time scale is
crucial. In many cases, the residence time in intermediates approaches the fs
regime characteristic of TS structures and the distinction becomes fuzzy.
Moreover, for a real multidimensional PES, the non-reactive nuclear motions
can entropically lock the system even though there is no well in the energy
landscape. And the presence of shallow wells in the energy landscape does
not guarantee that trajectories will visit such wells.
Transition State and Spectroscopy
Various techniques have been advanced to probe transition states more di-
rectly, especially for elementary reactions. Polanyi's analogy [14] of transition-
state spectroscopy, from “spectral wing emission”, to (Lorentz) collisional line
broadening studies, made earlier by A. Gallagher and others [see 12], set the
stage for the use of CW spectroscopic methods as a probe. (In this way, only
about one part in a million of the population can be detected.) Emission, ab-Ahmed H. Zewail 289
sorption, scattering and electron photodetachment are some of the novel
methods presented for such time-integrated spectroscopies, and the groups
of Jim Kinsey, Philip Brooks and Bob Curl, Benoit Soep and Curt Wittig, Dan
Neumark, and others, have made important contributions to this area of re-
search. The key idea was to obtain, as Kinsey [15] puts it, short-time dynamics
_from long-time experiments, (The renaissance of wave packet dynamics in spec-
troscopy, pioneered by Rick Heller, will be highlighted in Section 114 and
TN6). With these spectroscopies in a CW mode, a distribution of spectral fre-
quencies provides the clue to the desired information regarding the distribu-
tion of the TS over successive configurations and potential energies. Recently,
this subject has been reviewed by Polanyi and the author and details of these
contributions are given therein [12], and also in [26]
(2) Arrow of Time
In over a century of development, time resolution in chemistry and biology
has witnessed major strides, which are highlighted in Fig. 8 [16]. As men-
tioned before, the Arrhenius equation (1889) for the speed of a chemical
reaction gave information about the time scale of rates, and Eyring and
Michael Polanyi’s (1931) microscopic theoretical description made chemists
think of the atomic motions through the transition state and on the vibratio-
nal time scale. But the focus naturally had to be on what could be measured
in those days, namely the slow rates of reactions. Systematic studies of reac-
tion velocities were hardly undertaken before the middle of the 19th century;
in 1850 Ludwig Wilhelmy reported the first quantitative rate measurement,
the hydrolysis ofa solution of sucrose to glucose and fructose [8, Laidler]. In
1901, the first Nobel Prize for chemistry was awarded to van't Hoff for, among
other contributions, the theoretical expressions (chemical dynamics) which
were precursors to the important work of Arrhenius on rates. Arrhenius too
received the Prize in 1903 for his work on electrolytic theory of dissociation.
‘A major advance in experiments involving sub-second time resolution was
made with flow tubes in 1923 by H. Hartridge and F. J. W. Roughton for solu-
tion reactions. Two reactants were mixed in a flow tube, and the reaction
products were observed at different distances. Knowing the speed of the flow,
one could translate this into time, on a scale of tens of milliseconds. Such
measurements of non-radiative processes were a real advance in view of the
fact that they were probing the “invisible”, in contrast with radiative glows
which were seen by the naked eye and measured using phosphoroscopes.
Then came the stopped-flow method (B. Chance, 1940) that reached the milli-
second scale. The stopped-flow method is still used today in biological kinetics.
Around 1950, a stride forward for time resolution in chemistry came about
when Manfred Eigen in Germany and R. G. W. Norrish and George Porter in
England developed techniques reaching the microsecond time scale [17]. For
this contribution, Eigen and Norrish & Porter shared the 1967 Nobel Prize.
The method of flash photolysis was developed by Norrish and Porter a few
years after World War Il, using electronics developed at the time. They pro-290 Chemistry 1999
ARROW OF TIME
Chemistry & Biology Vee
LOE
Felacetion Mysthod
Flash Photons
Stogped Flew
Flow Metiod
Kingdes
Phuiens
Figure 8. Arrow of Time in chemistry and biology, some steps over a century of development (see
text). (Ref. BI]
duced an intense burst of light and created radicals in the sample, and, using
other light, they recorded the spectra of these radicals. They achieved kine-
on this time scale and observed some relatively stable intermediates.
Before the turn of the 20th century, it was known that electrical sparks and
Kerr cell shutters could have response times as short as ten nanoseconds. In
an ingenious experiment, Abraham and Lemoine (1899) [18] in France de-
monstrated that the Kerr response of carbon disulfide was faster than ten na-
noseconds; it has now been measured to be about two picoseconds (with
femtosecond response) [18]. They used an electrical pulse which produced a
spark and simultaneously activated a Kerr shutter. Light from the spark was
collimated through a variable-delay path and through the Kerr cell (polarizer,
tiAhmed H. Zewail 291
CS, cell and analyzer). The rotation of the analyzer indicated the presence of
birefringence in the cell for short optical delays; this birefringence disappear-
ed for pathlengths greater than several meters, reflecting the total op-
tical/electrical response time of 2.5 ns. In this landmark “pump-probe” expe-
riment, they demonstrated in 1899 the importance of synchronization. The
setting of time delays was achieved by varying the light path. Bloembergen
[18] has recently given a historical perspective of short-pulse generation and
Shapiro has reviewed the early developments, including mechanical, streak,
spark, stroboscope, and other high-speed photography methods [see Refs. 2
& 18]. As pointed out in these references [18], flash photolysis utilized the
above approach but one of the flashes was made very strong to generate high
concentrations of free radicals and hence their utility in chemical and spec-
troscopic applications.
Eigen developed “the relaxation method”, which reached the microsecond
and close to the nanosecond scale. By disturbing the equilibrium of a solu-
tion by either a heat jump, a pressure jump, or an electric field, the system
shifts from equilibrium. This is the point of time zero. Then the system equi-
librates, and its kinetics can be followed. (At about the same time, shock-tube
methods were used to provide kinetics on similar time scales.) Eigen called
these reactions "immeasurably fast” in his Nobel lecture. There was a feeling
that this time resolution was the fastest that could be measured or that need-
ed to be measured for relevance to chemistry (Section IV). The invention of
the laser has changed the picture.
Shortly after the realization of the first (ruby) laser by Maiman (1960), the
generation of giant and short pulses became possible: nanoseconds by Q-
switching (Hellwarth, 1961) and picoseconds (De Maria, et al 1966) by mode-
locking (1964). Sub-picosecond pulses from dye lasers (Schafer & Sorokin,
1966) were obtained in 1974 by Chuck Shank and Eric Ippen at Bell Labs,
and in 1987 a six £8 pulse was achieved [19]. In 1991, with the discovery of
femtosecond pulse generation from solidstate Ti-sapphire lasers by Sibbett
and colleagues [19], dye lasers were rapidly replaced and femtosecond pulse
generation became a standard laboratory tool; the state-of-the-art [19], once
8 fs, is currently about 4 fs and made it into the Guinness Book of World
Records (Douwe Wiersma’s group [19]). The tunability is mastered using
continuum generation [19, Alfano & Shapiro] and optical parametric ampli-
fication.
In the late sixties and in the seventies, ps resolution made it possible to
study non-radiative processes, a major detour from the studies of conventional
radiative processes to infer the non-radiative ones. As a beginning student, 1
recall the exciting reports of the photophysical rates of internal conversion
and biological studies by Peter Rentzepis [20]; the first ps study of chemical
reactions (and orientational relaxations) in solutions by Ken Eisensthal [21];
the direct measurement of the rates of intersystem crossing by Robin
Hochstrasser [22]; and the novel approach for measurement of ps vibrational
relaxations (in the ground state of molecules) in liquids by Wolfgang Kaiser
and colleagues [23]. The groups of Shank and Ippen [19] have made impor-292 Chemistry 1999
tant contributions to the development of dye lasers and their applications in
the ps and into the fs regime. Other studies of chemical and biological non-
radiative processes followed on the ps time scale, the scale coined by G. N.
Lewis as the “jiffy” - the time needed for a photon to travel 1 cm, or 33 pico-
seconds [24].
At about the same time in the sixties, molecular-beam studies of reactions
were being developed, and, although I was not initially a member of this com-
munity, beams later became part of our effort in femtochemistry. Molecular
collisions occur on a shorter time scale than a ps and real time studies were
not possible at the time. Crossed molecular beams and chemiluminescence
techniques provided new approaches for examining the dynamics of single
collisions using the postattributes of the event, the reaction products. The
contributions by Dudley Herschbach, Yuan Lee and John Polanyi [14, 25]
were acknowledged by the 1986 Nobel Prize. From state and angular distri-
butions of products, information about the dynamics of the collision was
deduced and compared with theoretical calculations of the PES and with mo-
lecular dynamics simulations; the goal was to find self-consistency and to
deduce an estimate of the lifetime of the collision complex. Crossed molecu-
lar beam-laser studies have probed dynamics via careful analyses of product
internal energy (vibrational and rotational) distributions and steady-state
alignment and orientation of products. The contributions to this important
area are highlighted in the article by Dick Zare and Dick Bernstein [25] and
in the book by Raphy Levine and Bernstein [9]. An overview of femtoche-
mistry (as of 1988) in connection with these other areas is given in a feature
article [26] by Zewail and Bernstein.
I. FEMTOCHEMISTRY: DEVELOPMENT OF THE FIELD
The development of the field is highlighted in this section, from the early
years of studying coherence to the birth of femtochemistry and the explosion
of research, or, as the report of Ref. 5 puts it, “... the revolution in chemistry and
adjacent sciences.” On the way, there were conceptual and experimental prob-
lems to overcome and many members of our Caltech group have made the
successful evolution possible. The review article published in the Journal of
Physical Chemistry [B14] names their contributions in the early stages of
development. Here, references are given to the work and explicitly in the
Figure Captions.
(1) The Early Years of Coherence
When I arrived in the US as a graduate student in 1969, nine years after the
invention of the first laser, I had no idea of what lasers were about. When ap-
pointed to the Caltech faculty as an assistant professor in 1976, I was not
thinking or dreaming of femtosecond time resolution. But I had the idea of
exploring coherence as a new concept in dynamics, intra-and inter-molecular.
This proved to be vital and fruitful.Ahmed H. Zewail 293
New Techniques for Molecules
The Caltech offer included start-up funds of $50,000 for capital equipment
($18K for shop services), an empty laboratory of two rooms and an office
next to it. A few months before moving to Caltech in May of 1976, I made the
decision not to begin with the type of picosecond research I was doing at
Berkeley as a postdoctoral fellow. Instead, the initial effort was focused on two
directions: (i) studies of coherence in disordered solids, and (ii) the devel-
opment of a new laser program for the studies of the phenomena of (optical)
coherence. Prior to the final move, I came down from Berkeley for several
visits in order to purchase the equipment and to outline the laboratories’
needs for electricity, water, gases and so on; my feeling was that of a man left
out in a desert with the challenge to make it fertile. We spent a significant
fraction of the $50K on setting up the apparatus for optical detection of mag-
netic resonance (ODMR) to study disordered solids. My experience at Penn
with Professor Robin Hochstrasser and at Berkeley with Professor Charles
Harris was to culminate in these experiments. The key questions I had in
mind were: What is the nature of energy migration when a crystal is systema-
tically disordered? Is it coherent, incoherent or partially coherent? Is there a
relationship between optical and spin coherence?
While the ODMR apparatus was being built, I was thinking intensely about
new laser techniques to probe the coherence of optical transitions (so-called
optical coherence). I had the intuitive feeling that this area was rich and at
the time had in mind several issues which were outlined in my research pro-
posal to Caltech. Laser experiments were designed with objectives focused on
the same issues outlined in the proposal. First, in the work I published with
Charles Harris, and alone at Berkeley, the coherence probed had been that of
spin (triplet excitons) and I felt that we should directly probe the optical
coherence, i.e., the coherence between the excited electronic state and the
ground state, not that between two spin states of the same excited state [Fig.
9]. Second, I did not believe that the time scale of spin coherence was the same
as that of optical coherence. Later, I wrote a paper on the subject which was
published in the Journal of Chemical Physics (1979) with the title: Are the
homogeneous linewidths of spin resonance and optical transitions related?
I was convinced that essentially all molecular optical transitions in solids
are inhomogencous, that is, they do not reflect the true dynamics of a homo-
geneous ensemble, but rather the overlapping effects of sub-ensembles. This
was the key point outlined in my research proposal to Caltech. If the new set
of laser experiments at Caltech proved successful, we should be able to find
the answer to these important (to me) questions. However, the funds re-
maining were insufficient to realize our dream for the new experiments.
Fortunately, Spectra Physics, started in 1961 and now a huge laser company,
was proud of a new product and was interested in helping us demonstrate the
usefulness of one of the first single-mode dye lasers they produced, At
Caltech, we had the pump argon-ion laser, from Wilse Robinson's Laboratory,
and added the dye laser we obtained from Spectra Physics with the idea that
we would purchase it if the experiments were successful. David Evans of204 Chemistry 1999
‘Coherent Transients — phase control
[EXACT SUPERPOSITION
#leate ge Lore pete) + ott) ]
LASER
Lisa
65 5 7 3 7
Tine (8)
Figure 9. Molecular coherence and dephasing. Coherent transients from the superposition of
states, and control of pulse phase (x or 3) in multiple pulse experiments, the optical analog of
NMR spectroscopy. The photon echo of iodine gas was observed on the spontaneous emission
using the described pulse sequence. [Ref. B9, B13, B22, B23, 50]
Spectra Physics was instrumental in helping us achieve this goal. What was left
were the low-temperature cryostat and electronics. We could not afford a
real metal cryostat, so we custom-made a glass Dewar that could go down to
1.8 Kelvin. Most of the electronics were obtained on loan for several months.
To generate laser pulses we used switching methods developed at JILA (Joint
Institute for Laboratory Astrophysics) and IBM. The work at IBM by Dick
Brewer's group triggered our interest in using electro-optic switching
methods.
We succeeded in making the laser perform according to specifications and
began the first experiments with phenazine crystals. This system, phenazine,
has unique properties which I learned about in Robin Hochstrasser’s labora-
tory. At Berkeley I published a paper outlining the nature of coherence in
multidimensional systems and used phenazine as a prototype experimentalAhmed H. Zewail 295
system. Our first laser transient at Caltech was beautiful. Unfortunately, the
transient was from the electronics, not from the crystal, but we soon realized
this! We decided to abandon this particular system for a while and to try an
impurity crystal of pentacene in a host of terphenyl. We also decided to study
gases, and some success came our way.
With the theoretical knowledge acquired in handling coherence effects,
which requires expertise with density matrix formalism and its manipulation
in geometrical frames, I had a novel idea: we should be able to detect coheren-
ce on the incoherent emission at optical frequencies. From many discussions
with Alex Pines and Charles Harris at Berkeley, I knew the power of “adding”
pulses in NMR and ESR. This idea was successful and indeed we were able to
observe the photon echo on the spontancous emission [Fig. 9] using three
optical pulses. Only months after my arrival in May of 1976, we published our
first scientific paper from Caltech. This success gave us confidence in the ap-
proach and in our understanding of the principles of coherence and its prob-
ing in molecular systems. We applied it to larger molecules with success but
also encountered some disappointments. This work was followed by a variety
of extensions to studies in gases and solids and also in a home-made (from
glass) effusive molecular beam. The small group in our laboratory of optical
coherence became productive and we had an exciting time. Our group and
that of Douwe Wiersma in Holland were then the two most active in these
areas of chemical research.
In the meantime, the work on disorder in solids began to yield interesting
results and, surprisingly, we observed an unusual change in the degree of
energy transfer with concentration, which we published as evidence of Ander-
son localization, a hot topic in the 1970's. This was followed by detailed
studies of several systems and I wrote my first research proposal to the
National Science Foundation (NSF). The research was funded! The then pro-
gram director at NSF, Fred Stafford, was supportive of the effort and we have
maintained our support from the Division of Materials Research to this day;
the NSF’s Chemistry Division continues to support our research. Knowledge
of energy transfer was also helpful in another area. Terry Cole, a visiting
scholar at Caltech, and I initiated work on the studies of Luminescent Solar
Concentrators (LSC) using energy transfer between dyes as a key principle
This idea, too, was successful and funding for this research came in from
SERI and from ARCO. The work on LSC resulted, ultimately, in a patent
(with Sam Batchelder; issued in 1980) and in several publications. My re-
search group was rapidly expanding.
After a year-and-a-half at Caltech, I was pleased to learn that my colleagues
were considering my case for tenure. Tenure was granted a few months later,
two years after my arrival at the Institute, I was both appreciative and pleased.
We continued research in four areas: optical coherence phenomena and
dephasing; disorder in solids; picosecond spectroscopy; and LSC. In my own
department, some colleagues were not too excited about ‘this stuff of cohe-
rence and dephasing’ ~ thinking that it was not relevant to chemistry! Many che-
mists on the outside were also unsure what this was all about. In fact, a notab-296 Chemistry 1999
le chemist once said publicly at a conference I attended that coherence and
dephasing had nothing to do with chemistry! On the physics side, I was in-
vited to numerous conferences, including one in which the Nobel Laureate
Willis Lamb asked me to have dinner to discuss our research. This was a spe-
cial experience!
I was not convinced by these doubts and my faith helped us to continue
along with the development. The concept of coherence turned out to be fun-
damental in femtochemistry, and it is now well accepted that coherence is a
key process in the probing and controlling of molecular dynamics. With the
success we had with observations and studies of coherence in different
systems, I wrote an Account of Chemical Research article, published in 1980,
with the title: Optical Molecular Dephasing — Principles of and Probings by Coherent
Laser Spectroscopy. 1 felt that the nanosecond time scale we had mastered
should be extended to the picosecond time scale, but did not wish to repeat
the Berkeley experience with glass lasers. Fortunately, the design for the first
sub-picosecond dye laser was reported in 1974 and we decided to build one to
study the phenomena of coherence ~ but now on a shorter time scale.
Optical Analogue of NMR: Controlling the Phase
From the studies of optical transients, we learned that coherence can be
probed directly in real time in gases (and solids) and that incoherent decay
(eg. fluorescence) can be used to monitor such coherences provided that
the laser pulse(s) is capable of forming a coherent superposition of states. For
two states of a transition (say y,, and y,), the coherent state can be written as
Peoherent () =a) ¥, + (DY, (4)
where the coefficients, a(t) and b(t) contain in them the familiar quantum-
mechanical phase factors, exp(iE,t/h) and exp (iE,t/f), respectively [see
Fig. 9].
With pulse sequences, we could directly monitor the behavior of the en-
semble-averaged coefficients of Y -'P*, (a(t) b*(t)), which contain informa-
tion on the coherence decay time (optical T,); they are the off-diagonal ele-
ments of a density matrix, p, The term (a(t) a*(t)) is the population of state
y,, and represents the diagonal density-matrix element, p,,; (a(t) a*(t)) de-
cays with optical T,. We were thus able to demonstrate the power of the opti-
cal analogue of NMR pulse techniques in learning about coherence and the
origin of optical dephasing in molecular systems of interest to chemical dy-
namics. This advance changed the thinking of many with the recognition that
it was impossible to deduce T, and T, from measurements of the line width of
inhomogeneous transitions.
One feature of this work which later helped us in the study of molecular
reaction dynamics was the realization of the importance of the pulse phase
(shape) in studies of coherence. With the acousto-optic modulation techni-
ques we developed earlier, it became possible to make optical pulse sequen-
ces with well-defined phases. This development took us into the domain of se-
lective and prescribed pulse sequences which could then be used to enhanceAhmed H. Zewail 297
coherences or suppress them ~ the optical analogue of NMR multiple pulse spec-
troscopy. We published several papers on phase control and extended the ap-
plications to include photon locking. We were eager to extend these techni-
ques to the picosecond time domain in order to study solids, but, for several
reasons, our attention was diverted to gas-phase molecular dynamics.
By this time, our group's efforts were narrowing on two major areas. (Dick
Bernstein, who was on the Visiting Committee for our Division, hinted that T
was doing too much in too many areas!) The work on disorder and LSC was
gradually brought to completion. Picosecond spectroscopy of rotational dif
fusion and energy transfer in liquids were similarly handled. I felt that the
latter area of research was too crowded with too many scientists, a characte-
ristic I do not enjoy when venturing into a new area. I must add that I was not
too thrilled by the exponential (or near exponential) decays we were measur-
ing and by the lack of molecular information. Our effort began to emphasize
two directions: (i) the studies of coherence and dynamics of isolated molecu-
les in supersonic beams and (ii) the development of the optical analogue of
NMR spectroscopy. The low-temperature facility was put to use to study the
dephasing and polarization of highly vibrationally-excited molecules in the
ground state. Coherence in chemical dynamics was occupying my thinking,
and I made a detour in the applications that turned out to be significant.
(2) The Marriage with Molecular Beams
The Bell Labs design for the dye laser (passively mode-locked, CW, and cavi-
ty dumped) was too restrictive for our use and, even though we published
several papers on studies in the condensed phase with 0.6 ps resolution, we
decided to change to a new system. The synchronously pumped mode-locked
(CW) dye laser allowed for tunability and also for photon-counting detection
techniques. The power of single-photon-counting became apparent, and a
new laser system, a synchronously pumped, cavity dumped dye laser was con-
structed for studies of gas-phase molecular dynamics, but now with the bene-
fit of all the expertise we had gained from building the first system used for
studies in the condensed phase and for probing the torsional of DNA.
Stimulated by the work on coherence, and now with the availability of pi-
cosecond pulses, I thought of an interesting problem relating to the question
of intramolecular vs. intermolecular dephasing. In large, isolated molecules
(as opposed to diatomics), there are the so-called heat bath modes which can
be a sink for the energy. The question arose: Could these bath modes in
isolated large molecules dephase the optically-excited initial state in the same
way that phonons of a crystal (or collisions in gases) do? This problem has
some roots in the question of state preparation, and I was familiar with its re-
lationship to the description of radiationless transitions through the work of
Joshua Jortner, Wilse Robinson and others. There was much theoretical ac-
tivity about dephasing, but I felt that they were standard extensions and they
did not allow for surprises. We decided on a new direction for the studies of
coherence in a supersonic molecular beam.298 Chemistry 1999
Rick Smalley came to Caltech in May of 1980 and gave a talk entitled
“Vibrational Relaxation in Jet-Cooled Polyatomics”, He spoke about his ex-
citing work on the naphthalene spectra. From the line width in the excitation
spectra, he inferred the “relaxation time”, At the time, the work by Don Levy,
Lennard Wharton, and Smalley on GW (or nanosecond) laser excitation of
molecules in supersonic jets was providing new ways to examine the spectros-
copy of molecules and van der Waals complexes. Listening to Rick and being
biased by the idea of coherence, I became convinced that the way to monitor
the homogeneous dynamics was not through the apparent width but by using
coherent laser techniques. This was further kindled by the need for direct
measurement of energy redistribution rates; we were encouraged by Charlie
Parmenter, after he had reported on a chemical timing method using colli-
sions as a “clock” to infer the rate of energy redistribution. The first “real” su-
personic molecular beam was huge. We did not know much about this kind of
technology. However, in a relatively short time, it was designed and built from
scratch, thanks to the effort of one graduate student who must have con-
sumed kilos of coffee! The molecular beam and picosecond system were in-
terfaced with the nontrivial addition of a spectrometer to resolve fluorescen-
ce in frequency and time. This was crucial to much of the work to come.
The Anthracene Discovery: A Paradigm Shift
Our goal in the beginning was to directly measure the rate of IVR (Intramo-
lecular Vibrational-cnergy Redistribution), expecting to see a decrease with
time (exponential decay) in the population of the initiallyexcited vibrational
state and to possibly see a rise in population in the state after the redistribu-
tion, thus obtaining T, directly. What we saw in these large systems was con-
trary to the popular wisdom and unexpected. The population during IVR was
oscillating coherently back and forth [Fig. 10] with well-defined period(s)
and phases! We were very excited because the results revealed the significan-
ce of coherence at its best in a complex molecular system, now isolated in a
molecular beam, with many degrees of freedom. I knew this would receive at-
tention and skepticism. We had to be thorough in our experimental tests of
the observation and three of us went to the laboratory to see how robust the
observation was. We published a Communication in the Journal of Chemical
Physics (1981). Earlier there had been attempts by another group to observe
such a “quantum coherence effect” in large molecules, but the observation
turned out to be due to an artifact. Some scientists in the field were skeptical
of our new observation and the theorists argued that the molecule is too big
to see such quantum coherence effects among the vibrational states. Further-
more, it was argued that rotational effects should wash out such an observa-
tion.
We followed the initial publication with several others and the effect be-
came even more pronounced with shorter time resolution. Physicists appre-
ciated the new results, We published a Physics Review Letter on the nature of
non-chaotic motion in isolated systems, and Nico Bloembergen and I wrote a
review (1984) on the relevance to laser-selective chemistry [B21]. We andAhmed H. Zewail 299
other groups subsequently showed the prevalence of this phenomenon in
large molecules. As is often the case in science, after the facts and in retro-
spect, the phenomenon was clear and was soon accepted; to some it even be-
came obvious! Looking back, this novel and unexpected observation was a
paradigm shift of critical importance, for a number of reasons:
First, the observation was the first to clearly show the presence of “quantum
coherence effect” in isolated complex chemical systems and only among se-
lected vibrational states. In other words, out of the expected chaotic motion in
the vibrational and rotational phase space we could see ordered and coherent mo-
tion despite the presence of numerous vibrational degrees of freedom (from
S, and §, states). This point was theoretically appreciated by only a few sci
tists. In fact, at one point Stuart Rice and I drafted a paper on the subject,
thinking of clarifying the point. At the time, researchers in high-resolution
spectroscopy were observing complex spectra and attributing this complexity
to chaotic vibrational motion in the molecule. Stuart and I argued that spec-
tral complexity does not mean chaos, and the anthracene experiment was a
clear demonstration in real time.
Second, the observation demonstrated that coherence had not previously
been detected in complex systems, not because of its absence but due to the
inability to devise a proper probe. Detection of total absorption or total emis-
sion (at all wavelengths) from molecules gives a non-selective window on the
dynamics and in this way coherence cannot be detected. This was a key point
for the success of the anthracene experiment for which both time and wave-
length were resolved and correlated. For all subsequent work on wave packet
dynamics, nuclear motion in chemical reactions, and femtochemistry, this
concept of “window probing” was essential. The concept was further elucidat-
ed by resolving the phase character. By probing at two different wavelengths,
we found that the quantum oscillations exhibit identical periods, but were
phase shifted by exactly 180° (i.e., they are out-ofphase) [Fig. 10]. The two
wavelengths resolved were those corresponding to emission of the initial
vibrational state and to that of the vibrational state to which the population
goes by IVR. Thus, if “total detection” was invoked, the in-phase and outof-
phase oscillations will add up to cancel each other and coherence would
have remained undiscovered.
Third, observation of phase coherent dynamics gave us a new dimension.
The phase shift indicates a true transfer of population, in contrast with con-
ventional quantum beats, and by analyzing the phases we could understand
the nature of IVR: “concerted”, i.e. going at the same time to all states, or “non-
concerted” i.e., going in a sequential redistribution of vibrational energy. We
could also obtain the time scale and the effect of molecular rotations on
coherence of the vibrational motion.
Fourth, the observation illustrated the importance of the “preparation of
nonstationary states” in molecules. This issue was of fundamental importan-
ce in radiationless transition theories involving multiple electronic states, and
experiments by Jan Kommandeur and by Doug McDonald have shown this in-
terstate coupling. The question of interest to us was: what nuclear states do we300 Chemistry 1999
IVR- Coherent and Dissipative Regimes
Coherent
MR
Clossicol-2
Noo
a >
Qvontuen-2
> + § a ib
6 &
‘Qvontum-N fe
Joy > tani) j
&
cc 4G
Time (nsec)
Figure 10. Dynamics of IVR, intramolecular vibrational energy redistribution. The coherent, re-
stricted and the dissipative regimes. Note the exact in-phase and out-ofphase oscillatory behavior
between the vibrational states of the system (anthracene in a molecular beam). The theory for
classical and quantum pictures (to the left) has been discussed in detail in the literature. [Ref.
51]; See also ref B20, B21.
prepare in the isolated molecule on a single surface? The anthracene experi-
ment taught us that a coherent source spanning the stationary states can pre-
pare a non-stationary state which evolves with time. Moreover, we can prepa-
re molecules in-phase at time zero to observe the subsequent coherent dyna-
mics. This concept indicates that the description, in terms of Schradinger’s
molecular stationary states, is not cardinal and that the time-dependent pic-
ture is real and directly relevant to dynamics. Most textbooks describe dyna-
mics in terms of stationary states and it took some time for this concept of a
time-dependent description to be appreciated. I recently found a theoretical
article by Roy Gordon published in the 1960's touching on similar issues. In
femtochemistry, the concept of time-domain dynamics is what describes ele-
mentary motions.
Fifth, by directly probing coherence and its extent in isolated, complex mo-
lecular systems we advanced some concepts regarding the nature of IVR and
its regions. We divided the regions of IVR into three basic ones: no IVR, re-
stricted IVR, and dissipative IVR. We also established that the IVR picture of one
vibrational state coupled to a continuum of vibrational levels is not adequate.
Instead it is a multi-tier coupling among vibrational states.Ahmed H. Zewail 301
This work and its implications were published in two series of papers and
reviewed in two book chapters.
The Successful 036 Laboratory
The laboratory known as 036 was in the sub-basement of Noyes. In thi
laboratory, the initial work on IVR was followed by fruitful applications span-
ning (i) studies of IVR in other systems, (ii) radiationless transitions, and
reaction rates of a variety of systems. One of our first studies of reactions on
the ps time scale, isomerization of stilbene, was stimulated by discussion with
Robin Hochstrasser about his work on stilbene vapor at room temperature.
He felt that if we could resolve the low-frequency modes in the molecular
beam, we would derive a great deal of information on the torsional potential.
We resolved these torsional modes. Furthermore, we decided to study the
rates as a function of energy and in the process found the barrier for twisting
around the double bond and observed coherent IVR in reactions, the first
such observation. Even now, stilbene remains a member of our molecular fa-
mily and continued studies have been pursued by us and others, also on the
fs time scale.
The following list highlights some of the work [Fig. 11] done in this initial
period from 1981 to 1983: (1) IVR in anthracene and stilbene; (2) trans-cis
isomerization of stilbene; (3) quantum beats and radiationless transitions in
pyrazine; (4) intramolecular hydrogen bonding in methyl salicylate; (5) in-
tramolecular electron transfer in donor-bridge-acceptor systems; (6) IVR and
dissociation of intermolecular hydrogen-bonded complexes; and (7) isome-
rization of diphenylbutadiene and styrene.
Over the years, in the same laboratory (036 Noyes), members of our group
have made new extensions covering the following topics: isomerization in iso-
lated molecules ys. in bulk solutions; nonchaotic multilevel vibrational ener-
gy flow; mode-specific IVR in large molecules; [VR dynamics in alky-anthra-
cenes; isotope effects on isomerization of stilbene; charge transfer and
exciton dynamics in isolated bianthracene; isotope effects on the intramole-
cular dephasing and molecular states of pyrazine; IVR dynamics in alkylanili-
nes (the "ring + tail” system); mode-specific (non-RRKM) dynamics of stil-
bene-are gas vdW complexes; solvation effects on intramolecular charge
transfer; IVR dynamics in p-difluorobenzene and p-luorotoluene (real time
vs. chemical timing); IVR dynamics in deuterated anthracenes; dynamics of
interstate coupling in chromyl chloride; dynamics of IVR and vibrational pre-
dissociation in anthracene-Ar, (n = 1,2,3); structural effects on the dynamics
of IVR and isomerization in stilbenes; and dynamics of IVR and vibrational
predissociation in n-hexane solvated transstilbene. The research resulted in a
series of publications.
Changing A Dogma: Development of RCS
The success with the anthracene experiment made us ask a similar question,
but now regarding the coherent rotational motion of isolated, complex mole-
cules. There were theories around which discarded its possibility because of= Chemistry 1999
Figure 11. Some examples of studies in the 036 Laboratory. The isomerization of stilbene, in-
tramolecular electron transfer, and solvation in clusters are examples of the studies made in the
early 1980's (see text); IQ = isoquinoline. [Ref. B3, 52]
the general belief that Coriolis interactions, anharmonicity and other inter-
actions would destroy the coherence. We worked out the theoretical implica-
tions and the results suggested possibly another surprise: if we could align the
molecules with a polarized picosecond pulse and probe (polarization-selective)
the rotating molecules, we should be able to observe rotational recurrences
which would give the full period of rotations of the isolated (large) molecule.
Classically, it is as if the molecule rotated back to its initial configuration. This
rotation period gives the moment of inertia and, since the masses of the
atoms are known, we can deduce distances, and hence obtain information on
molecular structures of very large molecules.
Indeed the recurrences in stilbene were observed with high precision, and
its molecular structure deduced. Coherence in rotational motion was clearlyAhmed H. Zewail 303
evident and could be probed in a manner similar to what we had done with
vibrational coherence. The approach was again met with some skepticism re-
garding its generality as a molecular structure technique. However, it is now
accepted by many as a powerful Dopplerfree technique; more than 120 struc-
tures have been studied this way. The method [Fig. 12] is termed “Rotational
Coherence Spectroscopy (RCS)” and is successfully used in many laboratories.
Some book chapters and review articles have been published on the subject
[see Bibliography].
Out of this first marriage between ultrafast lasers and molecular beams
came the developments and concepts discussed above. We were now poised
to study molecules and reactions with even shorter time resolution. We could
study their vibrational and rotational dynamics and align (“orient”) them by
controlling time.
Figure 12. Rotational Coherence Spectroscopy (RCS), the concept and first experimental obser~
vation made in a beam of transstilbene. [Ref. B7, B13, 53]304 Chemistry 1999
(3) The Transition to the Sub-Picosecond Regime
By the early 1980's, our laser time resolution for studying molecules in super-
sonic jets was 15 picoseconds and detection was made using a microchannel
plate (~ 40 ps). With this resolution, we had already studied reactions such as
the isomerization (twisting) of stilbene, charge-transfer, and intra- and inter-
molecular proton and hydrogen atom transfer. How could we improve the
time resolution and study, in a general way, the elementary steps of reactions?
The only approach I knew of was to use two pulses, one to “pump” and the
other to “probe”. Unlike liquid-state studies, where the approach was proven
successful, in this case, the density of a molecular beam is very low. Further-
more, it was not clear how to establish the zero of time in sitwin the molecu-
lar beam and how to avoid temporal broadening due to propagation effects.
A New Beam Machine: Pump-Probe Mass Spectrometry
I thought we should build a second generation beam apparatus to house a
time-of-flight mass spectrometer, From the physics literature, it was clear that
single-atom detection could be observed using ionization techniques with
lasers, and such detection had already been successful with nanosecond
lasers. Unlike many nanosecond studies, we should propagate the two pico-
second (and later femtosecond) pulses in the same direction, otherwise we
would lose the ultrashort time resolution! The same beam machine was
equipped with optics for laser-induced fluorescence detection. We began a
new direction of research in a separate laboratory of our group. The new
beam was built and integrated with two independently tunable dye lasers.
This proved to be a precursor to the femtochemistry work as this taught us to
master pump-probe picosecond and sub-picosecond experiments on chemi-
cal reactions. In this same laboratory, we studied with a resolution of a few ps:
(1) dissociation reactions; (2) ground-state, overtone-initiated reactions; (3)
van der Waals reactions and others [Fig. 13]. We wrote a series of papers on
state-to-state microcanonical rates, k(E), and addressed theoretical consequences
and deviations from the statistical regime.
It was in two of these systems (reactions of NCNO and ketene) that we
found that the statistical phase-space theory, although successful in describ-
ing productstate distributions, failed in describing the microcanonical rates
k(E) as a function of energy. Moreover, we made careful studies of the effect
of rotational population on k(E), and the effect was dramatic near the thres-
hold. Rudy Marcus, stimulated by these studies of k(E), applied variational
RRKM theory and we published some papers in a collaborative effort. The
key point here is that the TS “moves” to different (shorter) distances along
the reaction coordinate at different energies; the cardinal definition is relax-
ed [sce above, Section II]. In another system (H,O,), we studied the ground-
state (“thermal”) reaction for the first time in real time by initiating the reac-
tion with direct excitation of the overtones of the OH stretch vibration [Fig
13]. The coupling between theory and experiment stimulated my interest inAhmed H. Zewail 305
Ground State, Local-Mode Reactions
Reaction Rates of NCNO as a Function of Energy
vo!
sxid"
cio em S00
FS
Eanes é
°
16808 WWouo 17800 1740017800 T7800
Vibrational Energy, E (wavenumbers)
ye(E) (/seconds)
Figure 13. Microcanonical Rates of Reactions: (top) Ground-state reaction of HO, initiated by
loca-mode excitation; (bottom) the dissociation of NCNO as a function of energy, showing the
breakdown of conventional phase-space theory at energies above threshold. (Ref. Bl, 54]
the nature of transition states which generally live for less than a picosecond.
The thirst for even shorter time resolution became real!
‘The First ICN Experiment: Sub-picosecond Resolution
In the early 1980's, the technology of pulse compression became available
and we ordered, from Spectra Physics, a pulse compressor ~ a fiber optic ar-
rangement to reduce the laser pulse width to sub-picosecond. The company
indicated that it would take them several months to build one, and that the
only one available was at Purdue University in the laboratory of Professor
Duane Smith, one of the first two graduate students I had at Caltech. I men-
tioned to Duane my excitement about the experiment, which was intended to
directly monitor the elementary bond breakage in a molecule, and asked if it
was possible to borrow his compressor. The triatomic molecule ICN was
chosen because the CN radical could be conveniently monitored by laser-in-Ahmed H. Zewail 305
Ground State, Local-Mode Reactions
Reaction Rates of NCNO as a Function of Energy
vo!
sxid"
cio em S00
FS
Eanes é
°
16808 WWouo 17800 1740017800 T7800
Vibrational Energy, E (wavenumbers)
ye(E) (/seconds)
Figure 13. Microcanonical Rates of Reactions: (top) Ground-state reaction of HO, initiated by
loca-mode excitation; (bottom) the dissociation of NCNO as a function of energy, showing the
breakdown of conventional phase-space theory at energies above threshold. (Ref. Bl, 54]
the nature of transition states which generally live for less than a picosecond.
The thirst for even shorter time resolution became real!
‘The First ICN Experiment: Sub-picosecond Resolution
In the early 1980's, the technology of pulse compression became available
and we ordered, from Spectra Physics, a pulse compressor ~ a fiber optic ar-
rangement to reduce the laser pulse width to sub-picosecond. The company
indicated that it would take them several months to build one, and that the
only one available was at Purdue University in the laboratory of Professor
Duane Smith, one of the first two graduate students I had at Caltech. I men-
tioned to Duane my excitement about the experiment, which was intended to
directly monitor the elementary bond breakage in a molecule, and asked if it
was possible to borrow his compressor. The triatomic molecule ICN was
chosen because the CN radical could be conveniently monitored by laser-in-