throbber
Tunable Solid-State Lasers
`
`
`PETER F. MOULTON, SENIOR MEMBER,
`
`tF,Et=.
`
`Invited Paper
`
`.soIi(l—sIuIc mcrliu /mrc
`Tultablc lasers‘ based on l()II~(/()p(.’(/
`enrwgerl from the /rlbora/ory into ('(.)IIlI7l(’I‘('i(I/ ]>mtlu<’/io/r
`in the
`/(IS! (lccude, and /raw! formrl sigiirficmtt rtpp/icalioiz in p/totontc
`sy.rtcn1.r rcsearc/r. In the future. I/icy ntay be a t't'i/icrtl L‘(1III])()Il(‘IIl
`of 5pat‘c*—/Jasea’ renmze .\‘(’It.\'lI1g systcrits, communirulions .s'_\'mvrts
`and laser /ncrlicine. In I/iris‘ article we /‘ct'r'cn= the limit" /1/z_\'.rr'('.s'
`ofttuirtblc solid~stme [[l.\‘L’l‘S and disc-tr.v.v I/ll.’ C‘/Ill/‘(I(‘!(’I'i.\'li£‘S of/lit’
`most significtz/it e$m})li.s'/zed rmrl enrarging .S’_V.\'l(‘/11.3‘.
`
`l.
`
`lNTRODUCTlON
`In the last decade tunable solid—state lasers have become
`an increasingly significant component of quantum elec-
`tronics. While they were first demonstrated in the l96t)‘s.
`they have only recently emerged from being a laboratory
`curiosity to playing key roles in variety of ClCClrt)~t)pli(.‘S
`systems. Compared to what, until recently. was the most
`widely used tunable system. the liquid—medium dye laser.
`solid~state laser media offer unlimited operating and "shelt“‘
`lifetimes along with the capability to store energy and thus
`generate high peak powers via Q—switching.
`in addition,
`for some systems one can obtain either a broader tuning
`range than a given dye or an extension to longer infrared
`wavelengths. Current applications of tunable $0ll(.l'5lZllC
`lasers include basic scientific investigations in atomic.
`molecular and solid-state spectroscopy, laboratory studies
`of new semiconductor and tiber—optic devices. generation
`of ultrashort pulses and amplification of such pulses to high
`peak powers. Future applications now under development
`include aircraft—and space—based remote-sensing lidars. sub-
`marine communication systems and laser medicine. This
`article will
`review the field by first providing a brief
`discussion of the physics of tunable solid—state lasers and
`then examining a variety of laser systems. as categorized
`by the laser—active ion.
`(To be precise. we will cover
`paramagnetic—ion lasers only; other solid—state laser media
`employing so-called color centers are considered as an
`entirely different class of laser systems.)
`
`It.
`
`PHYSICAL BACKGROUND
`
`Solid-state lasers operate on stimulated transitions be-
`tween electronic levels of ions (activators) contained in
`Manuscript received August
`lél. lW|; revised October 2‘),
`lltttl.
`The author is with Schwartz Elcctro-Optics, lnc.r C‘oncord. MA (H742.
`lF.F.E Log Number 9l(l(iS33.
`
`solid crystalline or glassy media (hosts). To date. activators
`of practical significance have been positively charged ions
`from the rare Earth or 3d transition—metal groups of the
`periodic table. While all solid—state lasers can operate over
`some range of wavelengths. and thus are tunable.
`it
`is
`common to consider “tunable" lasers as those capable of
`covering a wavelength range greater than several percent of
`the laser central wavelength.
`In the following we discuss
`some basic laser concepts and consider the mechanisms
`important for making a laser "tunable?
`
`A. Lirmwitltlr. Ct'0.rS .S'eCrion. and Lifclinre
`
`The physical characteristic determining tuning range is
`the linewidth of the laser transition. Two other quantities
`of general interest to laser design are the gain cross section
`and the lifetime of the upper laser level.
`If we establish. by optical pumping. a population density.
`N (per cm”). of ions in ripper laser level. and limit our
`discussion to four—level lasers where the population of ions
`in the lower level can be neglected. then the optical gain.
`G’.
`in a laser medium of length 1 (in cm) is given by the
`expression
`
`C.‘ : t‘X])((f(/\\),i\‘—/)
`
`where rr(/\) is the gain cross section in cmz. as a function
`of wavelength ,\. The linewidth of the laser transition is a
`measure of the range of wavelengths over which the cross
`section is large, and is usually given as the t‘ull~width at
`hall‘-mttxirnum. ie. the span between the halllgain points
`on either side of the peak wavelertgtlt. (Unfortunately.
`in
`the field of quantum electronics, wavelength and frequency
`are interchanged all
`to frequently. We will generally stay
`with wavelength as the convention for this paper when
`rcfcrring to laser radiation. except for cases where the use
`of frequency is common.)
`The population of the upper laser level. in the absence of
`laser action. decays at it rate determined by the combined
`effects of radiative (or. spontaneous) emission and other
`nonradiative processes considered below. in simple systems
`the rate is constant. which leads to an exponential decay
`oi’ the population after the pumping is turned off. and :1
`characteristic lifetime given by the inverse of the decay rate.
`Litictimes for levels used in solid—statc lasers are in the range
`
`llll l t\’~“)E'. l ‘)1’
`
`L "Sift llll
`
`‘-T2: W02 IEEE
`
`l’R(')(‘l3El)l.\l(iS 0|" THF lEEE. VOL .\’lt_
`
`.\’O 3.
`
`.\l.»\RCH [W3
`
`ASML 1018
`
`ASML 1018
`
`

`
`from 10‘8 to 10‘2 s. The upper-level lifetime is important
`in determining how much pumping power is required to
`generate the necessary population of ions for laser action.
`For a desired steady-state population, N, the pump rate, W,
`in photons/s is simply NT, where r is the lifetime.
`If the only way the upper laser level decays is by spon-
`taneous emission to the lower laser level, the relation (first
`developed by Einstein) between stimulated and spontaneous
`emission can be used to connect the gain cross section,
`lifetime and linewidth by the following expression:
`
`am = F(,\)/\5 - tsmficr / F(A)AdA]“
`where F(/\) is the measured spectral emission curve for
`the laser transition, in power per unit wavelength interval,
`nistherefractiveindexofthc|xaer¢1'!313Landcisthe
`speed of light. If F(A) can be fit to a simple function,
`such as a Gaussian distribution, the above equariorrcan be
`simplified, but the basic result of the equation is the product
`of the gain cross section and 7' is inversely proportional to
`the linewidth.
`
`large gain
`For tunable systems with large Iinewidtls,
`cross sections‘ are possible only if the lifetime is short.
`Conversely, ‘for constant lifaimes, the gain cross section
`is inversely‘-related.to the linewidth.
`If one assumes a
`fixed length of lasenmaterial andcertain level of gain
`needed-«for lasu operation, one can readily show that
`the purnping.—po~ver' required to obtain steady-state laser
`operation is proportional to the linewidth. These kinds of
`consideratiors led early investigators to search for solid-
`state laser materials with narrow; linewidths. As techniques
`for optical pumping improved, including the use of-another
`laser as the pump source, operation on systems with large
`linewidths became possible.
`
`B. Ions in Host Crystals
`A solid—state laser host provides more than just a means
`for containment of the laser-active ions. In most systems,
`the environment -surrounding the ion serves -to modify
`the ion electronic states in a favorable way, by ‘making
`possible the interaction between electromagnetic radiation
`and the electronic states»-necessary for stimulated emission.
`In terms of quantum mechanics, the environment can induce
`between electronic states a dipole moment that would not
`exist if the ion was. in free space. For example, many of
`the laser transitions we will discuss below involve electrons
`that stay in the 3d state, and in free space the electric dipole
`moment for such transitions is parity forbidden, i.e., zero.
`Another function of the host crystal is to serve as a sink
`for excess energy produced by the pumping and lasing
`cycles of the solid-state system. Typically, the laser ions are
`excited by optical pumping, and if the pump source is either
`a gas discharge or tungsten lamp, much of the pump energy
`is absorbed into ion states higher in energy than that of the
`upper laser state. Decay of the excited ion from the higher
`states into the upper state must occur rapidly, accompanied
`by -a loss of ion energy, if the pumping process is to be
`effective. In a fourrlevel laser system even more energy
`
`MOULTON: TUNABLE SOLID-STATE LASERS
`
`_
`
`must be lost after the laser transition takes place, as the
`ion needs to rapidly decay from the lower laser level into
`the ground state. The host crystal creates a possibility for
`rapid loss of energy during the lasing cycle by allowing
`“nonradiative” transitions to occur between various states
`of the‘ ion.
`1
`In a nonradiative transition, some of the ion energy is
`convened to mechanical energy in the form of vibrations
`of the atoms in the host medium, because of vibrationah
`electronic interaction, or coupling. A quanturmnechanical
`treatrnenl,-’of vibrations in a solid yields a large set of
`discrcw. vibrational states, which are represented by a set
`of quasipanicles called phonons. Phonon energy is simply
`the vibrational frequency times 75,’ Ptanck’s constant; the
`properties of the host crystal (number and types of atoms
`in the unit cell,.the overall arrangement of atoms in the
`crystal structuee and the bonding strength between atoms)
`determine the number of distinct types of phonons and
`the distribution of phonon energies. All solid media have
`some characteristic maximum vibration frequency and thus
`a maximum phonon energy. The total number of phonons
`present
`in the host medium is related to the medium
`temperature. Every time a nonradiative transition occurs
`pbonons are created or, in more classical terms, the medium
`temperature increases. The rates at which nonradiative
`transitions take place cover a wide range and depend on a
`number of variables, including the properties of the active
`ion, the host crystal, the crystal temperature and the amount
`of energy given up into phonons.
`In many systems the
`rate is greater than 10"/s, and this is rapid enough for
`most laser pumping cycles. Even when the energy gap
`between levels is greater than the maximum phonon energy,
`there exist processes. in which multiple phonons can be
`created simultaneously, although the rate for such processes
`decreases rapidly as the number of phonons increase.
`While nonradiative transitions are crucial for promoting
`the decay of levels, they present a problem if they also cause
`decay- of the upper laser level. As we will discuss later,
`the presence of-_ nonradiative decay of the upper level can
`have a major effect on the performance of some,solid—state
`tunable lasers. ln a host crystal in which all the ions are in
`the same environment, the fluorescence quantum efficiency
`for a transition is the spontaneous emission rate divided by
`theqtotal transition rate. When all the decay is by photon
`emission the fluorescence quantum efficiency is unity.
`
`C. Broad Linewidths
`
`We first consider the mechanisms giving rise to the finite
`gain linewidth of a solid-state laser. Unlike gas lasers, the
`Doppler effect due to the motion of the ions in the host is
`not of significance. Vibrational frequencies range up to 1013
`Hz, but the ions are constrained to tiny displacements 00'“
`m) by the strong forces holding the host medium together,
`and- the resultant maximum velocities and Doppler shifts
`are relatively small.
`The most fundamental cause of a finite linewidth is the
`finite lifetimes of the upper and lower laser levels, since
`the uncertainty relation requires that the linewidth of any
`349
`
`

`
`level is the inverse of the level lifetime. To invoke quantum
`mechanics again, we note that the wavcfunction for a given
`ion level has both an amplitude and a phase. in calcttlating
`the value to be used for the lifetime of the level we must
`
`consider processes which not only cause decay ottt of the
`level. bttt also processes which change the phase. For all
`but
`the lowest
`temperatures the latter occur at
`a much
`higher rate than the former, and are due to interactions
`with the phonons of the host medium.
`in many common
`fixed~wavelength solid—state lasers, the linewidth is related
`to finite level lifetimes; values at room temperature range
`from 100 CH2 and on up. Two well—known examples are
`ruby, with the transition—mctal active ion Cr" in the host
`A1303 (sapphire), and Nd:YAG with the rare—Earth active
`ion Nd“ in the garnet—structured host Y3Al_<O,3.
`Lifetime broadening is homogeneous, that is, all the ions
`in the host are subject to the same broadening mechanism.
`For disordered host materials, such as glasses, another type
`of broadening occurs because each ion is in a slightly
`different environment from another. Since the energy levels
`of the ion depend to some extent on the surroundings,
`a distribution of ion energy levels results. The enormous
`number of ions that
`interact with a typical optical beam
`effectively create a smooth distribution of energies, and
`transitions between levels have a broadened lineshape. This
`type of broadening is referred to as inhomogeneous. since
`the properties of ions vary from one to the other. Familiar
`systems with inhomogeneous broadening include various
`Nd3*~doped glasses used as active media for laser drivers
`in inertial~confinement—fusion experiments and another rare
`Earth ion, En”. doped in silica fibers and used as amplifiers
`for l550—nm. ftber—transmission systems.
`The mechanism giving rise to the large linewidths of
`nearly all broadly tunable solid—state lasers is different from
`the two mentioned above, and comes about because of
`even stronger interaction between ions and the host than
`previously described. We refer to Fig.
`1
`to illustrate the
`mechanism.
`In the figure we plot energy as a function
`of
`the distance between an individual
`laser-active ion
`and the surrounding ions of
`the host crystal,
`for
`two
`different electronic states (1 and 2) of the ion and for
`two different levels of interaction. Our energy plot
`is the
`combined electronic energy of the ion and vibrational (i.e..
`mechanical) energy of the host crystal, and is referred to
`a configuration—coordinatc diagram. The energy versus
`distance relation is in the form of a parabola, a quadratic
`approximation to the manner in which the atoms of the
`crystal react to a change in spacing. The host vibrations we
`referred to earlier occur about the lowest energy point of
`the parabola, the equilibrium position. For this discussion
`we consider a simple mode of vibration in which all
`the
`atoms surrounding the |ascr~active ion move in and out at
`the same time, the “breathing” tnode. (in real host crystals
`there are many different modes of vibration. with much
`greater complexity of motion than the breathing mode. and
`each has a particular interaction with the electronic states
`of the active ion.)
`When the active ion changes front state 1
`
`to state 2. the
`
`350
`
`ENERGY
`
`‘
`
`** DISTANCE ~'*>
`
`(b)
`(it)
`Fig. 1. Energy as a function of distance between active inn and
`surrounding ions of host crystal. for two different electronic states.
`I and 2. in (a). surrounding inns do not shift equilibrium positions
`when active ion changes electronic states. while in (b) a shit"!
`occurs.
`
`equilibrium position may not be affected at all. as indicated
`in Fig. 1(a). This is typical for the 4f energy states of rare-
`Earth ions such as Nd, which have weak interactions with
`their surroundings. On the other hand.
`if the surrounding
`atoms are affected by the active~ion electronic state‘ they
`may react by establishing a new equilibrium position. as
`shown itt Fig.
`l(b). The latter case leads to the broad
`lincwidths characteristic of most tunable solid-state lasers.
`as we illustrate in Fig. 2.
`to
`Cmsider first an electronic transition from state i
`state 2, which would result from the absorption of energy.
`An important point about
`the transition is that
`the time
`in which it
`takes place is much shorter than the inverse
`of the vibrational
`frequencies of the atoms in the host
`medium. As a result.
`the atoms are essentially "frozen"
`during the transition. Since the atoms are vibrating.
`there
`is in fact a distribution of possible positions for the atomic
`configuration at the point of an electronic transition. and as
`Fig. 2 indicates. this leads to a wide distribution of possible
`transition energies. and thus a broad linewidth. Note that it
`similar transition for the system represented in Fig. 1(a)
`would have an energy independent of the atomic position.
`as the two parabolas have exactly the same spatial function.
`Immediately after the upwards transition both the host
`crystal and the active ion are in higher-energy states. The
`first component to lose energy is the host crystal. and it does
`so when the highly localized vibrational energy around the
`ion is dissipated into the rest of the crystal. If we invoke the
`concept of phonons again, we would say that phonons local
`to the ion are “emitted” and travel away from the ion. After
`this happens the combined ion~crystal system reaches the
`“relaxed“ excited state shown in Fig. 2. in which the crystal
`is in equilibrium, while the ion is still
`in an electronic
`excited state.
`in the absence of laser action, the ion makes an electronic
`transition from state 2 to state 1 due to a combination
`oi’ spontaneous emission and nonradiative processes. With
`laser action, stimulated emission provides another channel
`
`l’ROL‘El;'DlN(iS OF Till." lFF.E. V01, 30, NO 3.
`
`.\i.»\R('H IWI
`
`

`
`
`
`"*DSTANCE "*>
`
`Fig. 2. Enery diagram similar to Fig. l(b) with details of up-
`ward (absorption) and downward (emission) electronic transitions
`shown. Absorption and emission are spread in energy due to motion
`about equilibrium position of ions. “Zero-phonon" energy,
`the
`purely electronic difference in energy between states l and 2 is
`indicated.
`
`for the transition. As in absorption, the transition linewidth
`is broad, but the distribution of emission energies peaks at a
`lower energy than that for absorption. After the downwards
`transition occurs,
`the crystal
`is again left
`in an excited
`vibrational state, and relaxation down to the equilibrium
`condition for state'1 progresses rapidly through phonon
`emission.
`
`Figure 2 illustrates how the cycle of absorption, phonon
`emission to the relaxed excited state, stimulated emission
`into the lower level and phonon emission into the relaxed
`lower state is equivalent to a four-level system of discrete
`electronic levels. The difference is that the energy levels
`are a combination of both electronic and vibrational states,
`which leads to the use of the term “vibronic” (vibrational-
`electronic) to describe the types of transitions involved.
`Otherterms applied are “phonon-terminated” and the less
`accurate “phonon—broadened,” the latter of which also ap-
`plies, as we have noted above, to transitions of the type
`shown in Fig. 1(a).
`In Fig. 2 wesliow a classical view of how smooth and
`broad lineshapes for absorption and emission result from
`the dilference in energy between two displaced parabolas.
`A more rigorous quantum-mechanical analysis of the line-
`shapes relies on a calculation of overlap integrals between
`wavefunctions of the phonons in states 1 and 2. This
`analysis predicts t:hat_the energies of vibronic transitions
`should appear as discrete values, separated by the energy
`of an individual phonon and clustered on either side of a
`“zero-phonon” energy. The latter is indicated in Fig. 2 and
`is the difference between the purely electronic energy of
`the two states. If there was no displacement ofparabolas
`the zero-phonon line would be the observed transition
`energy in absorption and emission. As the displacement
`between the parabolas increases,
`theory shows that
`the
`intensity of the zero~phonon transition decreases while that
`of vibronic transitions increases. In actuality the discrete
`vfbronic transitions predicted by quantum mechanics are
`rarely observed, for-several reasons. First, each vibronic
`
`MOULTONI TUNABLE SOLID-STATE LASERS
`
`transition has a large linewidth from lifetime broadening,
`and the individual vibronic lines often overlap enough
`to produce a broad, featureless overall lineshape. Second,
`typically many different phonons, rather than one, interact
`with the electronic transition, and the effect of the resul-
`tant distribution of phonon frequencies obscures individual
`vibronic lines.
`We have assembled, in Fig. 3, a composite of the absorp-
`tion and emission spectra for many of tunable solid—state
`laser systems we will discuss in more detail below.
`ln
`all cases we show the emission band associated with laser
`operation and the absorption band for the same electronic
`transition. For some systems we also include highenlying
`absorption bands from other electronic transitions. In the
`systems considered emission bands connected with the
`more energetic transitions are barely, if at all, observable,
`as the excitation energy decays rapidly via phonon emission
`into the lowest-lying electronic excited state.
`
`D. Energy Storage and Ezrractiorr
`We now consider some of the issues related to Q-
`switching of and short-pulse amplification by tunable solid-
`state lasers. One of the favorable attributes of solid-state
`lasers in general
`is their ability to generate high peak
`powersby the Q-switching process. in Q~switching,
`the
`laser material is excited by the pump source but laser action
`is prevented by maintaining the laser cavity in a high-loss
`(low Q) state. When a high level of excitation in the upper
`level is attained, the cavity is switched to a hign—Q state
`i.e., the cavity loss is reduced and the laser output rapidly
`builds up from noise to a point at which a good fraction
`of the energy stored in the laser medium as excitation is
`extracted. If the rate of. extraction is much faster than the
`rate at which excitation in the medium can build up again,
`the laser output falls to zero in a short time. (After the
`extraction occurs the gain in the medium becomes less than
`the loss in the laser cavity.) Thus a pulse is generated.
`Energy in the laser cavity in the absence of gain in the
`laser medium decays exponentially with a characteristic
`time -rc, related to the round-trip time for the cavity and
`the loss fraction on each round-trip. Typical values for T,
`are in the ns range and Q~switched output pulsewidths are
`some multiple of ’T¢. A crude estimate of the ratio of the
`peak_ power generated by Q-switching to the power under
`normal conditions is T/Tc. The long upper-state lifetimes
`for most solid-state laser materials lead to high Q—switched
`powers, in contrast to dye and some gas media, which have
`ns-range values for 7'.
`Let us consider how energy storage affects the generation
`of high Q-switched powers. If the gain cross section for a
`material is too high, it may be difficult to store much energy
`in the laser medium before the Q-switch changes to the low-
`loss state. Even‘ for switches with essentially infinite loss,
`such as rotating mirrors, stored energy can be lost through
`parasitic oscillations within the laser medium, Solid-state
`lasers are particularly prone to parasitics, given that
`the
`medium has a high refractive index. High indices make
`possible unwanted laser cavities formed by reflections off
`351
`
`

`
`laser media surfaces tall in the range 5-30 I/_/cnr". for Q-
`switched pulses of approximately 10-30 ns in duration. For
`Nd:YAG lasers, with an f‘}_.,., of 0.4 J/cmg. damage and
`thus energy extraction is not a major concern. but as we
`point out below, values of Em,
`for many tunable laser
`media are l0~l00;r this level, and energy extraction from
`Q-switched oscillators is damage-limited.
`For short~pulse amplifiers, as long as the output fiuence
`of the amplifier is well below Em the pulse amplification
`is linear and energy extraction is not an issue. However.
`most amplifiers are operated to extract as much possible
`stored energy as possible from the medium. and simple
`analysis shows that
`at high fraction of the stored energy
`can be obtained when the output lluence is at
`least
`twice
`Em‘. As for Q—switched systems. the high values of I-.‘..,,
`for many tunable media prevent efficient amplifier opera-
`tton.
`
`E. Exct'ted—State Absorption
`
`As a final part of our background discussion. we consider
`the issue of excited-state absorption tF.SA). a process that
`for many tunable solid—state lasers can have a profound
`effect on system performance. Consider the energy diagram
`of Fig. 4. in which we include electronic states above the
`upper energy state for a laser transition. (We have elim—
`inated the atomic displacement
`to simplify the diagram.)
`Light can be either amplified by stimulated emission. where
`the ion drops to a lower-energy state. or can be attenuated
`through ESA, in which the ion makes a transition from the
`upper laser level to the higher—lying level. The net optical
`gain in the system is simply the difference between the
`cross section for stimulated emission and the cross section
`for ESA.
`for
`than that
`is greater
`If the cross section for ES/\
`stimulated emission laser action is impossible. no matter
`how high the level of pumping. Even if
`the net gain
`is positive. ESA raises the threshold pump power
`for
`laser action and reduces the efficiency in converting pump
`power
`to laser output power. A variation on ES/\
`is
`sometimes observed, in which the pump light is absorbed by
`a transition originating on the upper laser level. The effect
`of “pump BSA" depends, among other things. on the level
`of population in the upper level under conditions of laser
`action. It is particularly detrimental when Q--switched oper-
`ation is tlesired. since the upper population can rise to high
`levels before the Q—switch is changed to the low~loss slate.
`Prediction of both the magnitude and wavelength depen-
`dence ot‘ ELSA cross sections is difficult for most all tunable
`solid-state lasers. Our ability to generate the energy curves
`shown in Fig. 2 from theory is too limited in accuracy for
`trztnsition—metal
`ions in crystals. Conventional absorption
`measurements only access transitions from the ground state
`of the ion to higher lying states, and are generally not
`indicative of the nature of transitions from the "rela\'ed“
`excited state of the upper laser level. ESA cross sections
`can be estimated by measuring the optical gain (or loss) in
`a pumped crystal as a function of wavelength.
`
`PROCEEDINGS OF THF IEEE. VOL. 50. NO. 3.
`
`.\l;\RL‘H |‘l‘l.‘
`
`Ti:sappltl\'e
`
`Alexamrite
`
`_
`
`,
`900
`
`.,
`Mon
`
`1303
`
`
`1500
`
`Co:MgFg
`
`‘v‘2_5Joo
`
`,
`
`X
`
`
`__
`ll‘
`-
`toinfiflurtrso ‘V2’U.£)“DH—"‘z2?tZl
`Wavelo.'tgth lnrnl
`
`
`zoo '
`500"
`km
`
`l j
`
`l i
`
`Fig. 3. Absorption (solid line) and emission (dashed line) line’
`shapes as a function of wavelength for several comhinuliotts oi"
`laser~active ions and host crystal. The sharp line in the alexnndrite
`emission curve is off~scale. The absorption and emission bands in
`forsterite result from both Cr” and Cr-4+ ions.
`
`the ends of the media or by internal “bounce” paths based
`on total internal reflection. Even in systems where parasitic
`oscillations have been eliminated. ASE (amplification in
`the medium of spontaneous emission from the upper laser
`level) serves to deplete upper-level stored energy,
`For amplification of short pulses, energy storage is also
`an issue.
`in a manner similar to (Q—switching,
`a short-
`pulsc amplifier is pumped up to a high excitation level and
`then the pulse to be amplified is passed through the laser
`medium. The amount of amplification is set by limits to
`the medium stored energy.
`In Ncl:YAG practical levels of energy storage are reduced
`by both parasitics and ASE to levels well below the
`theoretical maximum possible, when all of the Nd ions are
`pumped to the upper laser level. The relatively (compared
`to Nd:YAG) low gain cross sections of most tunable solid-
`statc lasers permit high levels of energy storage in both
`(._)~switching and amplification,
`in some cases close to the
`theoretical
`limit.
`
`Unfortunately, a competing issue for Q~switched sys~
`terms. the ability to extract energy, is a major problem for
`many tunable systems. An estimate of the energy tluence
`(.1/cmii)
`in El Q~switched laser cavity can be obtained
`by calculating, from the gain cross section,
`the saturation
`tluence f§_.,.,,
`tor the laser transition, given by:
`
`E5." = /tr:/(/\rI)
`
`is Planck’s constant. Note that the above expres-
`where /1
`sion is simply the energy of a laser photon divided by the
`gain cross section. The actual fluence in the laser cavity
`is some factor, proportional
`to the pumping ratio (pump
`energy/threshold pump energy) multiplied by EN...
`Typical laser—induced damage levels for coated optics and
`
`-3.t.:
`
`

`
`Pump
`excited-state ‘
`absorption
`I
`
`i Excited-state
`: absorption
`
`
`
`Laser transition
`
`E. _ _' Qfj
`
`Sitnplitied electronic energy-level diagram showing tran-
`Fin. -8.
`sitittnx ;t\\(iCill1t.’\i with laser action. excitcd~state absorption and
`pinup L-sciterl—statc absorption.
`
`lll. DlVAt.l-NT 'l‘RANstTioN METALS
`
`xi. }-Ii.rIorii'u/ 8rIr'/tgrvtirtri
`ill
`the remainder of the paper. we will discuss specific
`utnahlc lasers. For historical
`reasons, we will start our
`tliscttssion with divalent (i.e., doubly charged) laser ions
`ltom the 3d transitiommetal group of the periodic table,
`as
`they were the first
`tunable solid-state lasers to be
`demonstrated. Indeed. they were the first examples ofa laser
`with the capability for broad, continuous tuning. in l963,
`Johnson. Dietz, and Guggenheim [ii at Bell Laboratories
`reported laser operation in the i620-um wavelength region
`lrottt crystals of Ni3”—doped MgF3. cryogenically cooled
`and pumped by a llashlamp; shortly afterwards [2],
`laser
`operation from Co“-doped MgF; and '/snF3 at various
`wavelengths between l”/50 and 2160 nm was obtained.
`Several other host crystals for Ni“ and Co“ ions were
`reported {3} in 1966. along with laser operation from a
`Ni:MgF3 crystal pumped by a CW tungsten lamp. and the
`lirst
`laser operation from the V3‘ ion in MgF3, The latter
`system. oscillating at ll2() nm, was discussed in more detail
`in a latter publication [4].
`In all cases the lasers were operated with the crystals
`cooled well below room temperature, usually at
`temper-
`atures close to that of liquid nitrogen. (77K). There were
`several
`reasons for
`this, of which the most
`important
`was thought to be the fact that the majority of the laser
`transitions at room temperature had low fluorescence quan-
`tttm elliciencics. At reduced temperatures the nonradiative
`decay rates were smaller,
`if not vanishing, and pumping
`powers became reasonable. The requirement for cryogenic
`cooling and the lack of a commercial source for laser
`crystals were early barriers to the transfer of divalent
`transition—mctal
`lasers from the initial
`laboratory experi-
`ments to widespread use.
`in the latter half of the l960’s
`the demonstration oi‘ dye lasers produced a great interest in
`tunable lasers, but in liquid, rather than solid media.
`ironically. developments in dye-laser technology were
`partly responsible for
`a new investigation of divalent
`tr2tnsition—rnetal lasers. begun in the latter half of the l97()’s
`by the author and coworkers at MIT Lincoln Laboratory [5].
`The optical pumping requirements for CW operation of dye
`lasers were such that only another laser could generate the
`
`Mlll.'l.'l’()t’\‘:
`
`'I‘LlN/\l5[.[i .'3'(H.lD-S'l3\TFZ LASERS
`
`needed pumping intensity. The concept of a lascr—pumped
`laser became widely accepted in the scientific community,
`and soon not only CW but pulsed laser—pumped dye lasers
`were standard laboratory devices. Laser pumping of solid-
`state materials was also possible, and for tunable media
`was applied to color-center lasers as early as 1974 [6]. For
`lasers where cryogenic cooling was considered necessary,
`laser pumping greatly simplified the system design.
`In
`the initial studies of divalent
`transition metal systems,
`flowing cryogenic liquids were generally used to cool
`laser rods, which were pumped through the rod cylinder
`surface by lamps. With laser pumping along the direction
`of optical gain (longitudinal pumping), conduction cooling
`of the crystal
`in a simple sta1ic—liquid Dewar could he
`accomplished, Fortunately, both Ni?‘ and Co“ ions can be
`pumped by either the 1300- or 1060—nm lines of neodymium
`(Nd)-doped solid—state lasers, while the absorption bands of
`the \/3* ion allow pumping by an argon—ion gas laser, and
`thus laser—pumped laser studies of the divalent transition
`metals were straightforward.
`The results of the work at Lincoln Laboratory can be
`summarized as follows: Analysis of laser operation in
`various Nizhdoped crystals showed that ESA of the laser
`wavelengths and possibly the pump wavelengths was sig—
`nificant and limited both laser efficiency and the maximum
`temperature at which laser operation could be obtained.
`Even at
`low temperatures, because of ESA the tuning
`range of Ni“ lasers was more limited than the fluorescence
`emission data predicted. For systems with little or no
`change in quantum cfficieney from cryogenic to room
`temperature, such as Ni:Mg0, cryogenic cooling was

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket