Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Weak-field multiphoton femtosecond coherent control in the single-cycle regime

Open Access Open Access

Abstract

Weak-field coherent phase control of atomic non-resonant multiphoton excitation induced by shaped femtosecond pulses is studied theoretically in the single-cycle regime. The carrier-envelope phase (CEP) of the pulse, which in the multi-cycle regime does not play any control role, is shown here to be a new effective control parameter that its effect is highly sensitive to the spectral position of the ultrabroad spectrum. Rationally chosen position of the ultrabroadband spectrum coherently induces several groups of multiphoton transitions from the ground state to the excited state of the system: transitions involving only absorbed photons as well as Raman transitions involving both absorbed and emitted photons. The intra-group interference is controlled by the relative spectral phase of the different frequency components of the pulse, while the inter-group interference is controlled jointly by the CEP and the relative spectral phase. Specifically, non-resonant two- and three-photon excitation is studied in a simple model system within the perturbative frequency-domain framework. The developed intuition is then applied to weak-field multiphoton excitation of atomic cesium (Cs), where the simplified model is verified by non-perturbative numerical solution of the time-dependent Schrödinger equation. We expect this work to serve as a basis for a new line of femtosecond coherent control experiments.

©2011 Optical Society of America

1. Introduction

Femtosecond laser pulses are unique tools to control multiphoton excitation processes in atoms and molecules [131]. The control over the corresponding initial-to-final state-to-state transition probabilities is achieved by shaping the pulse [3234] to manipulate the interferences among the coherent manifold of state-to-state multiphoton quantum transitions photo-induced by the broadband ultrashort pulse. The pulse-shaping control knobs are generally the relative phase, amplitude, and polarization of different spectral components of the pulse. Here we focus on rational coherent control approach, where the femtosecond pulse shaping is based on the initial identification of the interfering transitions and their interference mechanism. Such an approach has been shown to be very powerful when the photoexcitation picture is available in the frequency domain [131], which is possible when a corresponding perturbative description of finite order is valid. The weak-field regime corresponds to a valid description by perturbation theory of the lowest non-vanishing order. For two-photon processes it is second-order perturbation theory, while for three-photon processes it is third-order perturbation theory.

Over the last decade weak-field multiphoton femtosecond coherent control has been demonstrated very successfully in many studies that have been conducted in the multi-cycle femtosecond regime [131], where the pulse duration of the (unshaped) transform-limited pulse is of many optical cycles. A natural way to further enhance the possible control is to enrich the variety of multiphoton transitions by a significant broadening of the exciting pulse spectrum up to an octave-spanning corresponding to the single-cycle regime [3541]. There, the transform-limited pulse duration is of about one optical cycle. The shaping of such ultrabroadband pulses has recently been reported [40, 41]. As opposed to the multi-cycle regime, where the electric field is effectively determined only by the instantaneous temporal frequency and temporal envelope profile, in the single-cycle regime it is also strongly affected by the global phase between them. This phase is referred to as the carrier-envelope phase (CEP) and is also equal to the global phase of the spectral field at the different pulse frequencies. During the process of coherent spectral broadening and creation of single-cycle pulses, the CEP is experimentally stabilized to preset values [3539]. The CEP has shown to play a significant role in multiphoton processes such as photoionization, high harmonics generation, and others [4249].

Here we study for the first time weak-field multiphoton coherent control using shaped ultrabroadband femtosecond pulses having a bandwidth of the single-cycle regime. We investigate phase control of atomic bound-bound non-resonant multiphoton transitions. The CEP of the pulse, which in the multi-cycle regime does not play any control role, is shown here to be a new effective control parameter that its effect is highly sensitive to the spectral position of the ultrabroad spectrum. Rationally-positioned ultrabroad spectrum generally induces coherently several groups of multiphoton transitions from the ground to the excited state of the system: transitions involving only absorbed photons as well as Raman transitions involving both absorbed and emitted photons. For example, for a two-photon excitation process, one group is of transitions involving two absorbed photons and the other group is of Raman transitions involving one absorbed photon and one emitted photon. The intra-group interference is controlled by the relative spectral phases of the different frequency components of the pulse, while the inter-group interference is controlled jointly by the CEP and the relative spectral phases.

The corresponding effect and control principle are as follows. For N-photon excitation process, the global phase associated with all the transitions of N absorbed photons is N×CEP, while the global phase associated with all the transitions of M absorbed photons and N−M emitted photons is (2M−N) ×CEP. Thus, overall, the CEP contributes a value of 2(M−N) ×CEP to the relative phase between the excitation amplitudes photo-induced by these different groups. This value is added then to the relative phase between the excitation amplitudes as determined by the intra-group interferences taking place separately within each of the transition groups. The higher is the order of the multiphoton excitation (N), the smaller is the minimal ultrabroad bandwidth allowing such a control. In the multi-cycle regime, the spectrum is not broad enough to photo-induce several types of N-photon transitions; it can induce only a single type. Thus, no CEP effect is possible in the multi-cycle regime.

Specifically, the present study includes the investigation of non-resonant two- and three-photon excitation in a simple model system within the perturbative frequency-domain framework. Then, the developed intuition is applied to weak-field multiphoton excitation in atomic cesium (Cs), where the simplified model is verified by non-perturbative numerical solution of the time-dependent Schrödinger equation.

2. Two-photon theoretical perturbative description

Consider the simple model system shown in Fig. 1(a) , with a weak-field non-resonant two-photon transition from the ground state |g of energy Eg to the excited state |f of energyEf, with the corresponding transition frequency of Ωfg=(EfEg)/. The |g and |f states are coupled via a manifold of intermediate states |n (not shown in Fig. 1(a)) of the proper symmetry that are not accessed resonantly by the excitation. For a weak excitation pulse, the final amplitude Af(2)of |f after the pulse is over is given by second-order time-dependent perturbation theory as [9, 10]

Af(2)ε2(t)exp(iΩfgt)dt,
with the temporal electric field of the pulse ε(t) given by
ε(t)=12E(t)exp[i(ω0t+ϕCE)]+c.c.,
where E(t) is the (Gaussian) temporal complex amplitude of the pulse, ω0 is the carrier frequency, ϕCE is the carrier-envelope phase (CEP), and c.c. stands for the complex conjugate term.

 figure: Fig. 1

Fig. 1 (a) Two-photon excitation scheme in the femtosecond single-cycle regime. Indicated are the two-photon absorption transitions and two-photon Raman transitions. (b) Ultrabroad spectrum of single-cycle pulse centered at Ωfg that provides two absorbed photons of frequencies around Ωfg/2 to the two-photon absorption transitions as well as two photons, one absorbed photon of frequency around 3Ωfg/2 and one emitted photon of frequency aroundΩfg/2, to Raman transitions. (c) Temporal electric field of the TL single-cycle pulse with ω0 =12500 cm−1 (800 nm) and different ϕCE. (d) Relative population of the excited state |f induced by the TL single-cycle pulse with different values of ϕCE and different carrier frequenciesω0 (normalized to the case ofϕCE=0).

Download Full Size | PDF

The spectral field, correlated to the temporal field via Fourier Transform, is given by ε˜(ω)=E˜(ω)exp(iϕCE)+c.c., with E˜(ω) being the spectral complex envelope. It is given by E˜(ω)=|E˜(ω)|exp[iΦrel(ω)], where |E˜(ω)| and Φrel(ω) being, respectively, the spectral amplitude and relative phase of frequencyω. For the (unshaped) transform-limited (TL) pulse, Φrel(ω)=0 for anyω.

Hence, Af(2) is given in the frequency domain as

Af(2)=Aabs(2)+ARam(2),
with
Aabs(2)=exp(i2ϕCE)μabs20E˜(ω)E˜(Ωfgω)dωARam(2)=μRam,12ΩfgE˜(ω)E˜(ωΩfg)dω+μRam,22ΩfgE˜(ωΩfg)E˜(ω)dω,
where the amplitudeAabs(2) is coherently contributed by all the |g|f two-photon absorption transitions provided by the pulse spectrum, while the amplitudeARam(2) is coherently contributed by all the possible |g|f two-photon Raman transitions provided by the pulse spectrum. These two groups of transitions are schematically illustrated in Fig. 1(a).

Two-photon absorption transitions involve two absorbed photons of frequencies ω and ω'=Ωfgω (satisfying ω,ω'<Ωfg), while two-photon Raman transitions involve an absorbed photon of frequency ω (satisfying ω>Ωfg) and an emitted photon of frequencyωΩfg. Absorption of a photon contributes a global phase of ϕCE to its corresponding transition amplitude, while a photon emission contributes a corresponding global phase of ϕCE. Hence, as seen in Eq. (1), the two-photon absorption amplitudeAabs(2) has a global phase of 2ϕCE, while the two-photon Raman amplitudeARam(2) has zero global phase. So, the CEP value strongly affects the corresponding inter-group interferences. Theμabs2, μRam,12 andμRam,22 are, respectively, the effective non-resonant couplings of the two-photon absorption transitions and different two-photon Raman transitions [9, 10]. These couplings of the different groups of transitions generally have different values since each two-photon transition group accesses a different range of intermediate excitation energies (see the dashed lines in Fig. 1(a)) and thus involves different off-resonance detunings from the various intermediate states |n.

The possibility to simultaneously induce different types of two-photon transitions is a key feature of the ultrabroadband pulses of the single-cycle regime considered here as compared to the broadband pulses of the multi-cycle regime. For pulses with stabilized CEP ϕCE, the amplitudes associated with the different transition groups have relative phase of Δϕ=2ϕCE that dictates the nature of the interference between them. The observable considered here as the control objective of the CEP and relative spectral phase is the final population Pf of state |f that is given byPf=|Af(2)|2.

3. Two-photon perturbative calculations for a model system

3.1 The model system

The model quantum system we consider here has two-photon transition frequency of Ωfg =12500 cm−1. When the ultrabroadband spectrum of the irradiating pulse is centered in the region of the two-photon transition frequency, i.e., ω0Ωfg, it effectively provides photon pairs to both types of the two-photon transition groups: (i) Each two-photon absorption transition involves two absorbed photons with frequencies ω1,ω2Ωfg/2 that are symmetrically located around Ωfg/2, such that ω1+ω2=Ωfg; (ii) Each two-photon Raman transition involves one absorbed photon of frequency ω13Ωfg/2 and one emitted photon of frequencyω2Ωfg/2, such that ω1ω2Ωfg. Hence, we consider in our study pulses having their Gaussian spectrum centered in the region of Ωfg=12500 cm−1. Their spectral intensity bandwidth (FWHM) is Δω=5425 cm−1 corresponding to TL pulse duration of 2.7 fs, which is equal to one optical cycle of ω0 =12500 cm−1 (800 nm). The spectrum of the 800-nm pulse and its TL temporal electric field with different ϕCE values are schematically shown in Figs. 1(b) and 1(c).

As formulated and described above, the exact amplitudes contributed by the different two-photon transition groups (Aabs(2) and ARam(2)) are determined, on one hand, by the values of the different two-photon couplings (μabs2, μRam,12,μRam,22) and, on the other hand, by the field of the pulse ε˜(ω) and its different characteristics. Here, we assume for simplicity thatμabs2=μRam,12=μRam,22.

3.2 Results for two-photon CEP and relative-phase control

First, we study in this section the dependence of the two-photon excitation probability on the value of the CEP for the unshaped TL pulses of different carrier frequencies. The corresponding dependence of Pf,TL on ϕCE, calculated using Eq. (1), is shown in Fig. 1(d) for different carrier frequenciesω0, i.e., different positions of the pulse spectrum. Each trace is normalized by the corresponding value of Pf,TL for zero CEP, i.e., Pf,TL(ϕCE=0). The observed dependence on ϕCE is explained very simply as follows. The field E˜(ω)of the TL pulse is real since Φrel(ω)=0. So, the intra-group interferences within Aabs,TL(2) and ARam,TL(2) are fully constructive and maximize the magnitude of each of them, while the inter-group relative phase between them is 2ϕCE with a corresponding periodicity of π. In the case of ϕCE=0 and π, the amplitudes Aabs,TL(2) and ARam,TL(2) have zero relative phase [see Eq. (1)] and, thus, their inter-group interference is fully constructive and maximizes Pf,TL(ϕCE) for the given ω0. On the other hand, for ϕCE=π/2 and 3π/2, Aabs,TL(2) and ARam,TL(2) have a relative phase of π and, thus, their inter-group interference is fully destructive and leads to the minimal value of Pf,TL(ϕCE) for the given ω0.

The modulation depth MTL obtained for a given ω0, defined as MTL=[max(Pf,TL(ϕCE))min(Pf,TL(ϕCE))]/[max(Pf,TL(ϕCE))+min(Pf,TL(ϕCE))] or, equivalently, MTL=[Pf,TL(ϕCE=0)Pf,TL(ϕCE=π/2)]/[Pf,TL(ϕCE=0)+Pf,TL(ϕCE=π/2)], is determined by the amplitude ratio rTL=|Aabs,TL(2)|/|ARam,TL(2)| via the relation MTL=2rTL/(1+rTL)2. The maximal value of MTL=1 is obtained when |Aabs,TL(2)|=|ARam,TL(2)|, while the minimal value of MTL=0 is obtained when |Aabs,TL(2)| =0 or |ARam,TL(2)| =0. Other values of MTL, in between 0 and 1, are obtained for non-equal non-zero magnitudes of |Aabs,TL(2)| and |ARam,TL(2)|. Here, as seen in Fig. 1(d), the modulation depth increases from close-to-zero value of MTL≈0.05 at ω0 = 14500 cm−1 (pink line) to MTL≈0.25 at ω0=13500 cm−1 (green line) to MTL≈0.80 at ω0=Ωfg=12500 cm−1 (red line) and to its maximal value of MTL=1 at ω0=11870 cm−1 (black line), which is red shifted from Ωfg. Then, it decreases again to a value of MTL=0.45 at ω0=10300 cm−1 (blue line).

This observed behavior can be understood as follows. Analyzing Eq. (1) for a given Ωfg, it can be seen that in the case of a Gaussian TL pulse the magnitude of |ARam,TL(2)| (involving frequency differences) depends on the couplings μRam,12 andμRam,22 and on the pulse spectral bandwidth Δω, however it is actually independent of the carrier frequency ω0. On the other hand, the magnitude of |Aabs,TL(2)| (involving frequency sums) is determined by the coupling μabs2 as well as by Δω and ω0. Thus, for a given set of two-photon couplings and a given spectral bandwidth Δω, a change in ω0 leads to change only in |Aabs,TL(2)|, while |ARam,TL(2)| remains unchanged, and thus to a change in the ratio rTL=|Aabs,TL(2)|/|ARam,TL(2)|. Specifically, as the pulse spectrum is shifted to lower frequencies (i.e., shifted to the red), |Aabs,TL(2)| and rTL increase. In our model, due to the assumption of μRam,12=μRam,22=μabs2, one obtains that |Aabs,TL(2)|<|ARam,TL(2)| and rTL<1 for ω0=Ωfg and, thus, for any ω0>Ωfg. Hence, MTL<1 for any ω012500 cm−1 and the MTL value increases as ω0 decreases. As ω0 further decreases below Ωfg and becomes more and more red-shifted from it, the values of rTL and MTL further increase (see above) up to the point of = 11870 cm−1, where |Aabs,TL(2)| and |ARam,TL(2)| become equal and, thus, rTL=1 and MTL=1. Further red-shifting of ω0 leads to a further increase of rTL beyond the value of one (rTL>1) and thus to a corresponding decrease of MTL below a value of one (MTL<1), as indeed is the case for ω0=10300 cm−1.

Quantitatively, based on Eq. (1), the carrier frequency ω0 at which |Aabs,TL(2)|=|ARam,TL(2)| is given by ω0,MTL=1=12[Ωfg+Ωfg28(Δω8ln2)2lnRcoupling], where Rcoupling=μRam,12+μRam,22μabs2 is the ratio between the sum of the two-photon Raman couplings and the two-photon absorption coupling. As seen, the value of ω0,MTL=1 and its shift from Ωfg generally depends on Ωfg, Δω and Rcoupling. When Rcoupling=1, one obtains ω0,MTL=1=Ωfg, i.e., the maximal modulation occurs when there is no shift of ω0 from Ωfg. Such a case is, for example, the case of μRam,12=μRam,22=12μabs2. When Rcoupling<1 or Rcoupling>1, one obtains, respectively, ω0,MTL=1>Ωfg or ω0,MTL=1<Ωfg, i.e., the maximal modulation occurs when there is, respectively, a blue or red shift of ω0 from Ωfg. The latter case corresponds to our model that includes the assumption of μRam,12=μRam,22=μabs2 corresponding to Rcoupling=2. Indeed, we have obtained the maximal modulation for ω0,MTL=1=11870 cm−1, which is red-shifted from Ωfg.

Next, as a benchmark case for the full phase control of the two-photon excitation, we consider shaping the pulses of different carrier frequencies and different CEPs with a relative spectral phase step of π, i.e., Φrel(ω)=0 for ω<ωstep and Φrel(ω)= π for ωωstep [10]. The position of the phase step ωstep is scanned across the ultrabroadband pulse spectrum. The corresponding results for different values of ω0 in the range of moderate red-detuning to moderate blue-detuning from Ωfg are shown in the different panels of Fig. 2 . For each value of ω0, the corresponding dependence of Pf on ωstep is shown for the cases of ϕCE=0 (blue line) and ϕCE=π/2 (red line) as well as for the case of non-stabilized CEP (black line). The latter results from an average over all the different traces of Pf, each corresponding to a different value of ϕCE within the range of 0 to 2π. All the traces presented in Fig. 2 for a given ω0 are normalized by the value of Pf excited by the corresponding TL pulse having non-stabilized CEP.

 figure: Fig. 2

Fig. 2 Coherent control of two-photon excitation in the model system: Excited state population induced by ultrabroadband pulses shaped with a spectral π phase step at different positions for several cases of carrier frequency ω0 and CEP ϕCE. Each trace is normalized by the population excited by the corresponding TL pulse having non-stabilized ϕCE.

Download Full Size | PDF

A feature to note of all the π-traces is that the asymptotic values of each trace, i.e., the Pf values when ωstep is located far outside the spectrum either to the red or blue, are equal one to the other. It results from the fact that these two asymptotes correspond to a constant spectral phase of, respectively, π or 0 that is applied across the whole spectrum and thus is globally added to ϕCE. So, effectively they are equivalent to two unshaped TL pulses with a difference of π in their CEP values (ϕCE and ϕCE+π) that, as discussed above and shown in Fig. 1(d), both yield the same value of Pf. It also worth noting that the difference between the asymptotic values of the π-traces with different CEP values at a given ω0 also appears in the CEP-dependence results presented in Fig. 1(d).

A noticeable characteristic of the π-traces corresponding to ω0=10300 cm−1, which are presented in Fig. 2(a), is the approximate symmetric double-well structure around ωstep=Ωfg/2 =6250 cm−1. This feature can be explained as follows, considering the case of a spectrum having ω0 that is strongly red-detuned from Ωfg. Similar to the red-detuned TL case discussed above, also here in such a case one obtains that |Aabs(2)|>|ARam(2)| and Pf|Aabs(2)|2, i.e., the main contribution to Pf originates from the two-photon absorption transitions. As previously observed for two-photon absorption in the multi-cycle regime [10], the π-trace corresponding to |Aabs(2)|2 has a double-well structure that is symmetric around ωstep=Ωfg/2 and its two minima of Pf=0 result from complete destructive interferences among the two-photon absorption transitions. This π-trace has no dependence on the CEP value. The deviations from such a perfect symmetry around Ωfg/2 as well as the CEP dependence, which are observed in the π-traces of ϕCE=0 and π/2 that are presented for ω0=10300 cm−1, result from the still-existing non-zero small values of ARam(2) that interferes with Aabs(2). These deviations from the perfect symmetry are of opposite direction (along the axis of ωstep) in these two CEP cases, hence they disappear in the corresponding π-trace of the non-stabilized CEP case.

A completely different characteristic is observed for the π-traces of ω0=13500 cm−1 that are presented in Fig. 2(d). It is the symmetric wide double-well structure that is located around ω0. This feature can be explained as follows, in the same spirit of the above explanation for the red-detuned case but considering here the case of a spectrum that is strongly blue-detuned from Ωfg. Here, one obtains that |Aabs(2)|<|ARam(2)| and Pf|ARam(2)|2, i.e., the main contribution to Pf originates from the two-photon Raman transitions. The π-trace corresponding to |ARam(2)|2 has a wide double-well structure that is symmetric around ω0 and its two minima of Pf=0 are located at ωstep=ω0±(Ωfg/2) (here, ω0±6250 cm−1), i.e., they are Ωfg apart one from the other along the axis of ωstep. They result from complete destructive interferences among the two-photon Raman transitions. Also here, this π-trace has no dependence on the CEP value. The CEP dependence observed for the different π-traces of ω0=13500 cm−1 is due to the still-existing non-zero small values of Aabs(2) that interferes with ARam(2).

As seen in Fig. 2 for ω0=11870 cm−1 [Fig. 2(b)] and 12500 cm−1 [Fig. 2(c)], the signatures of these large-detuning π-trace features also show up when the pulse spectrum is located in between these two extreme spectral detuning cases. However, the exact shape of the corresponding π-trace results from the overall detailed intra- and inter-group interferences of the phase-shaped Aabs(2) and ARam(2), where both of them are generally of significant magnitude.

Another general important feature of the single-cycle regime analyzed in the present work, i.e., when ω0 is in the region of Ωfg, is the fact that the final population Pf excited by the unshaped TL pulse can sometime be exceeded by a properly shaped pulse. This is as opposed to the multi-cycle regime, where the (CEP-independent) Pf value induced by the unshaped TL pulse can never be exceeded by any shaped pulse; At most, it can be matched. This single-cycle feature is illustrated in all the π-traces of ϕCE=π/2 (red lines) that are shown in Fig. 2, where, for example, the population excited by the shaped pulse of ωstep=ω0 is higher than the population excited by the unshaped TL pulse. As discussed above, the latter corresponds to the asymptotic value of the π-trace. For ϕCE=0 (blue lines), on the other hand, the relative population excited by the TL pulse is still the maximal one. Below we analyze in detail these cases.

In the case of the TL pulse of ϕCE, as also discussed above, the intra-group interferences within Aabs(2) and ARam(2) are fully constructive and thus maximize |Aabs(2)| and |ARam(2)|, while the relative phase between Aabs(2) and ARam(2) is CEP dependent. For ϕCE=π/2, this relative phase is equal to π and thus the inter-group interference between the two amplitudes is completely destructive. For ϕCE=0, the inter-group relative phase is zero and the corresponding interference is fully constructive.

In the single-cycle case of ωstep=ω0, when ω0 is in the region of Ωfg, all the effectively contributing two-photon absorption transitions involve only pairs of photons that their frequencies are located near Ωfg/2, for which the π phase step contributes zero phase. Thus, the intra-group interferences within Aabs(2) are fully constructive and thus maximize |Aabs(2)|, while the total phase of Aabs(2) is Φ(Aabs(2))=2ϕCE+0=2ϕCE [see Eq. (1)]. On the other hand, all the effectively contributing two-photon Raman transitions involve pairs of photons that their frequencies are located in different spectral regions with respect to ωstep and thus the π phase step contributes a phase of 0 to one of them and a phase of π to the other. Hence, also the intra-group interferences within ARam(2) are fully constructive and maximize |ARam(2)|, however the total phase of ARam(2) is actually Φ(ARam(2))=π [see Eq. (1)]. Hence, here, for ϕCE = π/2, the relative phase between Aabs(2) and ARam(2) is zero and thus the inter-group interference between the two amplitudes is completely constructive. For ϕCE=0, the inter-group relative phase is π and the corresponding interference is fully destructive.

So, overall, as seen in the results of Fig. 2, one indeed obtains Pf,TL(ϕCE=π/2)<Pf,shaped,ωstep=ω0(ϕCE=π/2) (red lines) and Pf,TL(ϕCE=0)>Pf,shaped,ωstep=ω0(ϕCE=0) (blue lines).

4. Three-photon theoretical perturbative description and calculations for a model system

4.1 Theoretical description

In this section we study in detail the case of the non-resonant three-photon excitation. In the weak-field regime the final amplitude Af(3)of the excited state |f is given by the third-order perturbation theory as

Af(3)ε3(t)exp(iΩfgt)dt.

In the frequency domain it is given by

Af(3)=Aabs(3)+ARam(3),
with
Aabs(3)=exp(i3ϕCE)μabs300E˜(ω1)E˜(ω2)E˜(Ωfgω1ω2)dω1dω2
ARam(3)=exp(iϕCE)[μRam,1300E˜(ω1)E˜(ω2)E˜(ω1+ω2Ωfg)dω1dω2+μRam,2300E˜(ω1+ω2Ωfg)E˜(ω1)E˜(ω2)dω1dω2+μRam,3300E˜(ω1)E˜(ω1+ω2Ωfg)E˜(ω2)dω1dω2],
where the amplitudeAabs(3) is coherently contributed by all the |g|f three-photon absorption transitions provided by the pulse spectrum, while the amplitudeARam(3) is coherently contributed by all the possible |g|f three-photon Raman transitions provided by the pulse spectrum. These two groups of transitions are schematically illustrated in Fig. 3(a) .

 figure: Fig. 3

Fig. 3 (a) Three-photon excitation scheme in the femtosecond single-cycle regime. Indicated are the three-photon absorption transitions and three-photon Raman transitions. (b) Ultrabroad spectrum of the single-cycle pulse centered at Ωfg/2 provides three absorbed photons of frequencies around Ωfg/3 to the three-photon absorption transitions as well as three photons, two absorbed photon of frequency around 2Ωfg/3 and one emitted photon of frequency around Ωfg/3, to the Raman transitions. (c) Relative population of the excited state |f induced by the TL single-cycle pulse with different values of ϕCE and different carrier frequencies ω0 (normalized to the case of ϕCE= 0).

Download Full Size | PDF

Three-photon absorption transitions involve three absorbed photons of frequencies ω1, ω2 and ω3=Ωfgω1ω2 (satisfying ω1,ω2,ω3<Ωfg), while three-photon Raman transitions involve two absorbed photons of frequency ω1, ω2 (satisfying ω1+ω2>Ωfg) and an emitted photon of frequency ω1+ω2Ωfg. Hence, the global phases of the amplitudes Aabs(3) and ARam(3) are, respectively, 3ϕCE and 2ϕCEϕCE=ϕCE, resulting in a relative phase between them of Δϕ=2ϕCE. The μabs3,μRam,13,μRam,23 andμRam,33 are, respectively, the effective non-resonant couplings of the three-photon absorption transitions and different three-photon Raman transitions.

The observable considered here as the control objective of the CEP and relative spectral phase is the final population Pf of state |f that is given byPf=|Af(3)|2.

4.2 The model system

The model system we consider has three-photon transition frequency of Ωfg=25000 cm−1. In the three-photon excitation case, the rational choice of the excitation pulse spectrum is with a carrier frequency ω0 in the region of one-half of the three-photon transition frequency, i.e., ω0Ωfg/2=12500 cm−1. As illustrated in Fig. 3(b), such spectral position effectively provides significant intensity at frequencies around Ωfg/3 as well as around 2Ωfg/3. Hence, as desired for CEP control, it generally yields significant magnitude to both Aabs(3) and ARam(3). The three-photon absorption amplitude Aabs(3) results from transitions involving three photons of ω1,ω2,ω3Ωfg/3 with ω1+ω2+ω3=Ωfg, while the three-photon Raman amplitudeARam(3) results from transitions involving three photons of ω1,ω22Ωfg/3 and ω3Ωfg/3 with ω1+ω2ω3=Ωfg. For consistency, we use here for the three-photon case the same pulse of the two-photon case, i.e., a pulse of 5425-cm−1 bandwidth corresponding to a 2.7-fs TL pulse that is of a single cycle at ω0=12500 cm−1 (800 nm). Generally, it worth noting that the higher is the order of the multiphoton excitation, the smaller is the minimal bandwidth of the ultrabroad spectrum that is required for observing a CEP control effect. This is important for some experimental configurations, where the pulse-shaping implementation is somewhat easier with a reduced bandwidth. Regarding the effective three-photon couplings, similar to the two-photon excitation case, we assume here for simplicity that μabs3=μRam,13=μRam,23=μRam,33.

In order to avoid here an additional source of interferences, the excitation amplitude induced via one-photon transition at frequency Ωfg, which is described theoretically by the first-order perturbative term, is not included in our model. This is justified since, as seen in Fig. 3(b), this spectral component falls far outside the main spectral region, with a corresponding intensity that is about three orders of magnitude lower than the intensity at ω0=Ωfg/2. Thus, experimentally the spectral tail located in the region of Ωfg can easily be eliminated by proper amplitude shaping of the pulse [3234].

4.3 Results for three-photon CEP and relative phase control

The results for the phase control of the three-photon excitation are shown in Figs. 3(c) and 4 . The former presents the CEP control results for unshaped TL pulses, while the latter presents the results for the full phase control utilizing both the CEP and relative spectral phase. Conceptually and qualitatively, the three-photon results are very similar to the control results of the two-photon case [Figs. 1(d) and 2], hence in the following we will only describe in short some of the features of the three-photon results.

 figure: Fig. 4

Fig. 4 Coherent control of three-photon excitation in the model system: Excited state population induced by ultrabroadband pulses shaped with a spectral π phase step at different positions for several cases of carrier frequency ω0 and CEP ϕCE. Each trace is normalized by the population excited by the corresponding TL pulse having non-stabilized ϕCE.

Download Full Size | PDF

The dependence of the three-photon excited population Pf,TL on the ϕCE value for the unshaped TL pulses is shown in Fig. 3(c) for different carrier frequencies ω0. Each trace is normalized by the corresponding value of Pf,TL(ϕCE=0). With a TL pulse, the magnitudes of Aabs,TL(3) and ARam,TL(3) are maximal due to fully constructive intra-group interferences within each of them, and the inter-group relative phase between them is 2ϕCE having a corresponding periodicity of π. With ϕCE=0 and π, Aabs,TL(3) and ARam,TL(3) have zero relative phase [see Eq. (2)] and, thus, their inter-group interference is constructive leading to the maximal Pf,TL(ϕCE) for the given ω0. With ϕCE=π/2 and 3π/2, Aabs,TL(3) and ARam,TL(3) have a relative phase of π and, thus, their inter-group interference is destructive and Pf,TL(ϕCE) is the minimal one for the givenω0.

In terms of the modulation depth MTL, also here, a spectrum that is largely red or blue shifted from Ωfg/2=12500 cm−1 leads to a magnitude |Aabs,TL(3)| that is much larger or much smaller, respectively, than |ARam,TL(3)|. Such detuned spectrum supports effectively only one group out of the two groups of three-photon transitions. The resulting ratio rTL=|Aabs,TL(3)|/|ARam,TL(3)| is, respectively, either much smaller or much larger than 1, and thus the modulation depth MTL in both cases is small [see blue and green lines in Fig. 3(c) forω0=10500 and 13500 cm−1]. As seen in Fig. 3(c), the maximal modulation depth of MTL=1 and the corresponding minimal TL population of zero, obtained when |Aabs,TL(3)|=|ARam,TL(3)|, occur here at ω0=11750 cm−1 (black line), i.e., at a moderate red shift from Ωfg/2. The cause for this maximal-modulation red-shift effect is qualitatively similar to the one of the two-photon excitation case. It is the combination of the following facts: (i) rTL<1 for ω0=Ωfg/2 due to our model assumption of μabs3=μRam,13=μRam,23=μRam,33, and (ii) rTL increases as ω0 increases.

Figure 4 presents the results for the full phase control of the three-photon excitation using shaped pulses of different carrier frequencies and different CEPs applied with a relative spectral phase step of π. The different panels of Fig. 4 present the results for different values of ω0. For each value of ω0, the corresponding dependence of Pf on the phase step position ωstep is shown for the cases of ϕCE=0 (blue line) and ϕCE=π/2 (red line) as well as for the case of non-stabilized CEP (black line). All the π-traces presented for a given ω0 are normalized by the value of Pf excited by the corresponding TL pulse having non-stabilized CEP.

As in the two-photon case, the asymptotic values of each π-trace are equal one to the other. It results from the fact that the two asymptotes correspond to a constant spectral phase of 0 or π that is globally added to ϕCE and, thus, they correspond to two unshaped TL pulses with a difference of π in their CEP values. As shown in Fig. 3(c), such two TL pulses induce the same populationPf. The difference between the asymptotic values of the π-traces with different CEP values at a given ω0 also appears in the CEP-dependence results of Fig. 3(c).

When ω0 is far detuned to the red from Ωfg/2 [Fig. 4(a)], as in the TL case, the dominant contribution to the total excited amplitude Af(3) is from the absorption part Aabs(3). Then, the resulting π-trace is similar to the one obtained for the three-photon absorption case in the multi-cycle regime [10], with three zero-value minima located around two peaks of small magnitude. On the other hand, when ω0 is far detuned to the blue from Ωfg/2 [Fig. 4(d)], the dominant contribution to Af(3) comes from the Raman part ARam(3) leading to a π-trace that is of similar structure to the red-detuned case but much wider. As the CEP value affects only the inter-group interferences, when one of the group amplitudes Aabs(3) or ARam(3) dominates, the π-trace is expected to have only weak dependence on the CEP value. Indeed, as seen in Figs. 4(a) and 4(d), in these two highly-detuned cases there is only a small difference between the π-traces obtained with ϕCE=0 and π/2.

When ω0Ωfg/2 [Figs. 4(b) and 4(c)], the exact shape of the corresponding π-trace is determined by the overall detailed intra- and inter-group interferences of the phase-shaped Aabs(3) and ARam(3), where both of them are generally of significant magnitude. For every three-photon transition that contributes to the absorption or Raman amplitudes, the position of the π-step determines the relative phases of the three involved photons and, hence, the way this three-photon transition interferes with the other transitions of his group. All such intra-group interferences determine the magnitudes of Aabs(3) and ARam(3), and, together with the CEP, they also determine the phases of Aabs(3) and ARam(3). All in all, these resulting amplitudes yield the final outcome of Pf.

Last, it is important to note that also the three-photon results exhibit the general important feature of the single-cycle regime that, as opposed to the multi-cycle regime, the final population excited by the TL pulse can sometime be exceeded by a properly shaped pulse. This is illustrated in the π-traces of ϕCE=π/2 (red lines) presented in Figs. 4(a) and 4(b).

5. Non-perturbative calculations for cesium (Cs) atom

In this section, the control strategy presented above is applied to atomic Cs, demonstrating it for its two excited states |f15d3/2 and |f25d5/2 that are populated by two-photon excitation from the ground state |g6s. The corresponding two-photon couplings between the ground and excited states are provided by the manifold of p-states of Cs. As opposed to our model system analyzed above, in a real atom, depending on its specific structure, the effective two-photon couplings μabs2, μRam,12, and μRam,22 generally have different magnitude and might also have different sign. The Cs excitation scheme is shown in Fig. 5 , with examples of the two-photon absorption and Raman transitions. The results for the two-photon excitation of the Cs atom irradiated by weak unshaped (TL) and shaped ultraboradband pulses have been calculated by the numerical propagation of the time-dependent Schrödinger equation using the fourth-order Runge-Kutta method. The calculations have considered the atomic states 6s to 9s, 6p to 8p, and 5d to 7d, with all their fine-structure states. The data of the states and the corresponding transition dipole moments are based on Refs [5053].

 figure: Fig. 5

Fig. 5 Weak-field two-photon femtosecond coherent phase control of Cs in the single-cycle regime. Upper panel: Excitation scheme of Cs (fine structure splitting is not shown), with examples of two-photon absorption and Raman transitions. Panels (a) and (b): Relative population of the excited states 5d3/2 [panel (a)] and 5d5/2 [panel (b)] induced by TL single-cycle pulses of ω0 = 16667 cm−1 (Δω = 7350 cm−1) having different values of ϕCE. Panels (c) and (d): Relative population of the excited states 5d3/2 [panel (c)] and 5d5/2 [panel (d)] induced by such ultrabroadband pulses applied with ϕCE = 3.3 rad and shaped with a spectral π phase step at different positions. Indicated are the different state-to-state transitions associated with the different resonance-mediated enhancement peaks of the π-traces (see text).

Download Full Size | PDF

Similar to the model results presented above, the Cs results also show high degree of CEP control using unshaped TL pulses, with a high sensitivity to the value of the pulse carrier frequency ω0 as compared to the value of the two-photon transition frequency. The transition frequencies corresponding to the Cs states 5d3/2 and 5d5/2 are, respectively, Ωf1,g = 14499.3 cm−1 and Ωf2,g = 14596.8 cm−1. We have found that the most effective CEP control over the Cs two-photon excitation occurs with ω0 = 16667 cm−1 (600 nm), where the single-cycle TL pulse is of 2-fs duration and bandwidth of Δω=7350 cm−1. The corresponding results of the CEP control with such 600-nm single-cycle unshaped TL pulses are shown for the final populations of the Cs states 5d3/2 and 5d5/2 in, respectively, Figs. 5(a) and 5(b). As seen, there is a strong dependence of the final populations of both states on the ϕCE value, also here, with a periodicity of π. The maximal modulation depth of MTL=1 is obtained for the 5d5/2-state population [Fig. 5(b)], while a very high modulation depth of MTL=0.83 is obtained for the 5d3/2-state population [Fig. 5(a)]. As also seen in Figs. 5(a) and 5(b), the minimal CEP-dependent TL populations of both states 5d5/2 and 5d3/2 are obtained for ϕCE=0.16 rad and ϕCE=0.16+π=3.3 rad, indicating that at these CEP values the inter-group interference between the corresponding overall TL two-photon absorption and Raman amplitudes is the most destructive one. These CEP values are different from the corresponding model-system values of ϕCE=π/2 and 3π/2 due to the resonance-mediated nature [11, 20] of some of the two-photon transitions involved in the Cs excitation with the ultrabroadband pulses. Such resonance-mediated transitions do not exist in the non-resonant model excitation analyzed above.

The Cs results also exhibit effective full phase control using the CEP and relative spectral phase of the pulse. As an illustrative example, Figs. 5(c) and 5(d) show the results for, respectively, the 5d3/2 and 5d5/2populations that are controlled with the above ultrabroadband pulses (ω0=16667 cm−1, Δω=7350 cm−1) applied with CEP of ϕCE=3.3 rad and shaped with a π phase step. One main feature is that, as seen, setting the π-step position in the central spectral region of the pulse (i.e., ωstepω0) converts the above-mentioned destructive inter-group interference of the TL case into a constructive one and leads to an enhancement of the excited populations. Also this effect is indeed predicated by the above model-system results.

Additional prominent features of the π-traces presented in Figs. 5(c) and 5(d) are the several enhancement peaks appearing there. They also originate from the above-mentioned resonance-mediated nature [11, 20] of some of the two-photon transitions induced in the ultrabroadband excitation of Cs. When ωstep is set exactly at a spectral frequency that is equal to a transition frequency between the initial ground state (6s) and an intermediate state (np) or between an intermediate state (np) and the final state (5d3/2 or 5d5/2), the two-photon excitation probability is significantly enhanced. As previously explained [11], such an enhancement results from the interferences among two-photon transitions that are near-resonant with the intermediate state and their corresponding detunings are of different signs. These (intra-group) interferences are of destructive nature when induced by the TL pulse, while they become to be constructive when induced by a pulse shaped with a properly-positioned π phase step. The different state-to-state resonant transitions leading to the different resonance-mediated enhancement peaks in the π-traces of the 5d3/2 and 5d5/2 populations are indicated in Figs. 5(c) and 5(d). It worth noting that the 5d3/2 and 5d5/2 populations are not equally enhanced even when the same intermediate state is involved. For example, the resonance-mediated enhancement of the 5d3/2 population is either larger or smaller than the enhancement of the 5d5/2 population when the involved intermediate state is either 6p or 7p, respectively. The corresponding reason is the difference in the effective two-photon couplings associated with the different excitation pathways, which in the above example are 6s-6p-5d and 6s-7p-5d.

6. Conclusions

To summarize, we have shown theoretically the power of weak-field multiphoton femtosecond coherent control in the single-cycle regime. In this regime, for a rationally chosen pulse spectrum, the CEP of the pulse serves as a powerful control parameter in addition to the relative spectral phase. For weak-field N-photon excitation process, the ultrabroadband pulse induces several groups of initial-to-final N-photon transitions, where each group corresponds to a different combination of the number of absorbed photons (M) and the number of emitted photons (N-M). The intra-group interference is phase controlled by the relative spectral phase, while the inter-group interference is phase controlled jointly by the CEP and relative spectral phase. Specifically, here, we have revealed and illustrated the corresponding control mechanism in non-resonant two- and three-photon excitation of a simple model system using frequency-domain perturbative analysis. Then, the phase control mechanism of the single-cycle regime has been confirmed also for weak-field two-photon excitation of atomic cesium (Cs) studied by non-perturbative numerical solution of the time-dependent Schrödinger equation. The same mechanism generally applies also to multiphoton excitations of higher order. We expect this work to serve as a basis for a new line of femtosecond coherent control experiments, most prominently for selective control in excitation scenarios that simultaneously involve several channels.

Acknowledgements

This research was supported by The Israel Science Foundation (Grant No. 1450/10) and by The James Franck Program in Laser Matter Interaction.

References and Links

1. D. J. Tannor, R. Kosloff, and S. A. Rice, “Coherent pulse sequence induced control of selectivity of reactions: Exact quantum mechanical calculations,” J. Chem. Phys. 85(10), 5805 (1986). [CrossRef]  

2. M. Shapiro, and P. Brumer, Principles of the quantum control of molecular processes (Wiley, New Jersey, 2003).

3. W. S. Warren, H. Rabitz, and D. Mahleh, “Coherent control of quantum dynamics: the dream is alive,” Science 259(5101), 1581–1589 (1993). [CrossRef]   [PubMed]  

4. R. J. Gordon and S. A. Rice, “Active control of the dynamics of atoms and molecules,” Annu. Rev. Phys. Chem. 48(1), 601–641 (1997). [CrossRef]   [PubMed]  

5. H. Rabitz, M. Motzkus, K. Kompa, and R. de. Vivie-Riedle, “Whither the future of controlling quantum phenomena?” Science 288(5467), 824–828 (2000). [CrossRef]   [PubMed]  

6. M. Dantus and V. V. Lozovoy, “Experimental coherent laser control of physicochemical processes,” Chem. Rev. 104(4), 1813–1859 (2004). [CrossRef]   [PubMed]  

7. Y. Silberberg, “Quantum coherent control for nonlinear spectroscopy and microscopy,” Annu. Rev. Phys. Chem. 60(1), 277–292 (2009) (and references therein). [CrossRef]  

8. P. Nuernberger, G. Vogt, T. Brixner, and G. Gerber, “Femtosecond quantum control of molecular dynamics in the condensed phase,” Phys. Chem. Chem. Phys. 9(20), 2470–2497 (2007) (and references therein). [CrossRef]   [PubMed]  

9. Y. Silberberg and D. Meshulach, “Coherent quantum control of two-photon transitions by a femtosecond laser pulse,” Nature 396(6708), 239–242 (1998). [CrossRef]  

10. D. Meshulach and Y. Silberberg, “Coherent quantum control of multiphoton transitions by shaped ultrashort optical pulses,” Phys. Rev. A 60(2), 1287–1292 (1999). [CrossRef]  

11. N. Dudovich, B. Dayan, Y. Silberberg, and S. M. G. Faeder, “Transform-limited pulses are not optimal for resonant multiphoton transitions,” Phys. Rev. Lett. 86(1), 47–50 (2001). [CrossRef]   [PubMed]  

12. N. Dudovich, D. Oron, and Y. Silberberg, “Single-pulse coherently controlled nonlinear Raman spectroscopy and microscopy,” Nature 418(6897), 512–514 (2002). [CrossRef]   [PubMed]  

13. H. U. Stauffer, J. B. Ballard, Z. Amitay, and S. R. Leone, “Simultaneous phase control of Li2 wave packets in two electronic states,” J. Chem. Phys. 116(3), 946 (2002). [CrossRef]  

14. B. Chatel, J. Degert, S. Stock, and B. Girard, “Competition between sequential and direct paths in a two-photon transition,” Phys. Rev. A 68(4), 041402 (2003). [CrossRef]  

15. B. Chatel, J. Degert, and B. Girard, “Role of quadratic and cubic spectral phases in ladder climbing with ultrashort pulses,” Phys. Rev. A 70(5), 053414 (2004). [CrossRef]  

16. A. Präkelt, M. Wollenhaupt, C. Sarpe-Tudoran, and T. Baumert, “Phase control of a two-photon transition with shaped femtosecond laser-pulse sequences,” Phys. Rev. A 70(6), 063407 (2004). [CrossRef]  

17. S. H. Lim, A. G. Caster, and S. R. Leone, “Single-pulse phase-control interferometric coherent anti-Stokes Raman scattering spectroscopy,” Phys. Rev. A 72(4), 041803 (2005). [CrossRef]  

18. X. Dai, E. W. Lerch, and S. R. Leone, “Coherent control through near-resonant Raman transitions,” Phys. Rev. A 73(2), 023404 (2006). [CrossRef]  

19. B. von Vacano and M. Motzkus, “Time-resolving molecular vibration for microanalytics: single laser beam nonlinear Raman spectroscopy in simulation and experiment,” Phys. Chem. Chem. Phys. 10(5), 681–691 (2008). [CrossRef]   [PubMed]  

20. A. Gandman, L. Chuntonov, L. Rybak, and Z. Amitay, “Coherent phase control of resonance-mediated (2 + 1) three-photon absorption,” Phys. Rev. A 75(3), 031401 (2007). [CrossRef]  

21. L. Chuntonov, L. Rybak, A. Gandman, and Z. Amitay, “Enhancement of intermediate-field two-photon absorption by rationally shaped femtosecond pulses,” Phys. Rev. A 77(2), 021403 (2008). [CrossRef]  

22. L. Chuntonov, L. Rybak, A. Gandman, and Z. Amitay, “Frequency-domain coherent control of femtosecond two-photon absorption: intermediate-field versus weak-field regime,” J. Phys. At. Mol. Opt. Phys. 41(3), 035504 (2008). [CrossRef]  

23. Z. Amitay, A. Gandman, L. Chuntonov, and L. Rybak, “Multichannel selective femtosecond coherent control based on symmetry properties,” Phys. Rev. Lett. 100(19), 193002 (2008). [CrossRef]   [PubMed]  

24. L. Rybak, L. Chuntonov, A. Gandman, N. Shakour, and Z. Amitay, “NIR femtosecond phase control of resonance-mediated generation of coherent UV radiation,” Opt. Express 16(26), 21738–21745 (2008). [CrossRef]   [PubMed]  

25. N. T. Form, B. J. Whitaker, and C. Meier, “Enhancing the probability of three-photon absorption in iodine through pulse shaping,” J. Phys. B 41(7), 074011 (2008). [CrossRef]  

26. N. Dudovich, T. Polack, A. Pe’er, and Y. Silberberg, “Simple route to strong-field coherent control,” Phys. Rev. Lett. 94(8), 083002 (2005). [CrossRef]   [PubMed]  

27. M. Wollenhaupt, A. Präkelt, C. Sarpe-Tudoran, D. Liese, T. Bayer, and T. Baumert, “Femtosecond strong-field quantum control with sinusoidally phase-modulated pulses,” Phys. Rev. A 73(6), 063409 (2006). [CrossRef]  

28. C. Trallero-Herrero, J. L. Cohen, and T. Weinacht, “Strong-field atomic phase matching,” Phys. Rev. Lett. 96(6), 063603 (2006). [CrossRef]   [PubMed]  

29. S. D. Clow, C. Trallero-Herrero, T. Bergeman, and T. Weinacht, “Strong field multiphoton inversion of a three-level system using shaped ultrafast laser pulses,” Phys. Rev. Lett. 100(23), 233603 (2008). [CrossRef]   [PubMed]  

30. A. Lindinger, C. Lupulescu, M. Plewicki, F. Vetter, A. Merli, S. M. Weber, and L. Wöste, “Isotope selective ionization by optimal control using shaped femtosecond laser pulses,” Phys. Rev. Lett. 93(3), 033001 (2004). [CrossRef]   [PubMed]  

31. W. Salzmann, U. Poschinger, R. Wester, M. Weidemüller, A. Merli, S. M. Weber, F. Sauer, M. Plewicki, F. Weise, A. M. Esparza, L. Wöste, and A. Lindinger, “Coherent control with shaped femtosecond laser pulses applied to ultracold molecules,” Phys. Rev. A 73(2), 023414 (2006). [CrossRef]  

32. A. M. Weiner, “Femtosecond pulse shaping using spatial light modulators,” Rev. Sci. Instrum. 71(5), 1929 (2000). [CrossRef]  

33. T. Brixner and G. Gerber, “Femtosecond polarization pulse shaping,” Opt. Lett. 26(8), 557–559 (2001). [CrossRef]  

34. M. Wollenhaupt, A. Assion, and T. Baumert, “Femtosecond laser pulses: linear properties, manipulation, generation and measurement”, Handbook of Lasers and Optics, ed. F. Träger, p. 937 (Springer, New York, 2007).

35. P. Dombi, A. Apolonski, Ch. Lemell, G. G. Paulus, M. Kakehata, R. Holzwarth, Th. Udem, K. Torizuka, J. Burgdörfer, T. W. Hänsch, and F. Krausz, “Direct measurement and analysis of the carrier-envelope phase in light pulses approaching the single-cycle regime,” N. J. Phys. 6, 39 (2004). [CrossRef]  

36. E. Matsubara, K. Yamane, T. Sekikawa, and M. Yamashita, “Generation of 2.6 fs optical pulses using induced-phase modulation in a gas-filled hollow fiber,” J. Opt. Soc. Am. B 24(4), 985 (2007). [CrossRef]  

37. S. Akturk, C. D'Amico, and A. Mysyrowicz, “Measuring ultrashort pulses in the single-cycle regime using frequency-resolved optical gating,” J. Opt. Soc. Am. B 25(6), A63 (2008). [CrossRef]  

38. S. Rausch, T. Binhammer, A. Harth, J. Kim, R. Ell, F. X. Kärtner, and U. Morgner, “Controlled waveforms on the single-cycle scale from a femtosecond oscillator,” Opt. Express 16(13), 9739–9745 (2008). [CrossRef]   [PubMed]  

39. A. L. Cavalieri, E. Goulielmakis, B. Horvath, W. Helml, M. Schultze, M. Fieß, V. Pervak, L. Veisz, V. S. Yakovlev, M. Uiberacker, A. Apolonski, F. Krausz, and R. Kienberger, “Intense 1.5-cycle near infrared laser waveforms and their use for the generation of ultra-broadband soft-x-ray harmonic continua,” N. J. Phys. 9(7), 242 (2007). [CrossRef]  

40. B. Xu, Y. Coello, V. V. Lozovoy, D. A. Harris, and M. Dantus, “Pulse shaping of octave spanning femtosecond laser pulses,” Opt. Express 14(22), 10939–10944 (2006). [CrossRef]   [PubMed]  

41. B. von Vacano, W. Wohlleben, and M. Motzkus, “Actively shaped supercontinuum from a photonic crystal fiber for nonlinear coherent microspectroscopy,” Opt. Lett. 31(3), 413–415 (2006). [CrossRef]   [PubMed]  

42. T. Nakajima and S. Watanabe, “Phase-dependent excitation and ionization in the multiphoton ionization regime,” Opt. Lett. 31(12), 1920–1922 (2006). [CrossRef]   [PubMed]  

43. T. Nakajima and S. Watanabe, “Effects of the carrier-envelope phase in the multiphoton ionization regime,” Phys. Rev. Lett. 96(21), 213001 (2006). [CrossRef]   [PubMed]  

44. A. Apolonski, P. Dombi, G. G. Paulus, M. Kakehata, R. Holzwarth, Th. Udem, Ch. Lemell, K. Torizuka, J. Burgdörfer, T. W. Hänsch, and F. Krausz, “Observation of light-phase-sensitive photoemission from a metal,” Phys. Rev. Lett. 92(7), 073902 (2004). [CrossRef]   [PubMed]  

45. A. J. Verhoef, A. Fernández, M. Lezius, K. O’Keeffe, M. Uiberacker, and F. Krausz, “Few-cycle carrier envelope phase-dependent stereo detection of electrons,” Opt. Lett. 31(23), 3520–3522 (2006). [CrossRef]   [PubMed]  

46. Y. Wu and X. Yang, “Carrier-envelope phase-dependent atomic coherence and quantum beats,” Phys. Rev. A 76(1), 013832 (2007). [CrossRef]  

47. S. Chelkowski and A. D. Bandrauk, “Sensitivity of spatial photoelectron distributions to the absolute phase of an ultrashort intense laser pulse,” Phys. Rev. A 65(6), 061802 (2002). [CrossRef]  

48. M. F. Kling, J. Rauschenberger, A. J. Verhoef, E. Hasovic, T. Uphues, D. B. Milosevic, H. G. Muller, and M. J. J. Vrakking, “Imaging of carrier-envelope phase effects in above-threshold ionization with intense few-cycle laser fields,” N. J. Phys. 10(2), 025024 (2008). [CrossRef]  

49. M. J. Abel, T. Pfeifer, A. Jullien, P. M. Nagel, M. J. Bell, D. M. Neumark, and S. R. Leone, “Carrier-envelope phase-dependent quantum interferences in multiphoton ionization,” J. Phys. At. Mol. Opt. Phys. 42(7), 075601 (2009). [CrossRef]  

50. N. I. S. T. Atomic Spectra Database, (NIST, Gaithersburg, MD) available at http://physics.nist.gov/asd.

51. S. A. Blundell, W. R. Johnson, and J. Sapirstein, “Relativistic all-order calculations of energies and matrix elements in cesium,” Phys. Rev. A 43(7), 3407–3418 (1991). [CrossRef]   [PubMed]  

52. M. S. Safronova, W. R. Johnson, and A. Derevianko, “Relativistic many-body calculations of energy levels, hyperfine constants, electric-dipole matrix elements, and static polarizabilities for alkali-metal atoms,” Phys. Rev. A 60(6), 4476–4487 (1999). [CrossRef]  

53. M. S. Safronova and C. W. Clark, “Inconsistencies between lifetime and polarizability measurements in Cs,” Phys. Rev. A 69(4), 040501 (2004). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 (a) Two-photon excitation scheme in the femtosecond single-cycle regime. Indicated are the two-photon absorption transitions and two-photon Raman transitions. (b) Ultrabroad spectrum of single-cycle pulse centered at Ω f g that provides two absorbed photons of frequencies around Ω f g / 2 to the two-photon absorption transitions as well as two photons, one absorbed photon of frequency around 3 Ω f g / 2 and one emitted photon of frequency around Ω f g / 2 , to Raman transitions. (c) Temporal electric field of the TL single-cycle pulse with ω 0 =12500 cm−1 (800 nm) and different ϕ C E . (d) Relative population of the excited state | f induced by the TL single-cycle pulse with different values of ϕ C E and different carrier frequencies ω 0 (normalized to the case of ϕ C E = 0).
Fig. 2
Fig. 2 Coherent control of two-photon excitation in the model system: Excited state population induced by ultrabroadband pulses shaped with a spectral π phase step at different positions for several cases of carrier frequency ω 0 and CEP ϕ C E . Each trace is normalized by the population excited by the corresponding TL pulse having non-stabilized ϕ C E .
Fig. 3
Fig. 3 (a) Three-photon excitation scheme in the femtosecond single-cycle regime. Indicated are the three-photon absorption transitions and three-photon Raman transitions. (b) Ultrabroad spectrum of the single-cycle pulse centered at Ω f g /2 provides three absorbed photons of frequencies around Ω f g /3 to the three-photon absorption transitions as well as three photons, two absorbed photon of frequency around 2 Ω f g /3 and one emitted photon of frequency around Ω f g /3, to the Raman transitions. (c) Relative population of the excited state | f induced by the TL single-cycle pulse with different values of ϕ C E and different carrier frequencies ω 0 (normalized to the case of ϕ C E = 0).
Fig. 4
Fig. 4 Coherent control of three-photon excitation in the model system: Excited state population induced by ultrabroadband pulses shaped with a spectral π phase step at different positions for several cases of carrier frequency ω 0 and CEP ϕ C E . Each trace is normalized by the population excited by the corresponding TL pulse having non-stabilized ϕ C E .
Fig. 5
Fig. 5 Weak-field two-photon femtosecond coherent phase control of Cs in the single-cycle regime. Upper panel: Excitation scheme of Cs (fine structure splitting is not shown), with examples of two-photon absorption and Raman transitions. Panels (a) and (b): Relative population of the excited states 5d 3 / 2 [panel (a)] and 5d 5 / 2 [panel (b)] induced by TL single-cycle pulses of ω 0 = 16667 cm−1 ( Δ ω = 7350 cm−1) having different values of ϕ C E . Panels (c) and (d): Relative population of the excited states 5d 3 / 2 [panel (c)] and 5d 5 / 2 [panel (d)] induced by such ultrabroadband pulses applied with ϕ C E = 3.3 rad and shaped with a spectral π phase step at different positions. Indicated are the different state-to-state transitions associated with the different resonance-mediated enhancement peaks of the π-traces (see text).

Equations (8)

Equations on this page are rendered with MathJax. Learn more.

A f ( 2 ) ε 2 ( t ) exp ( i Ω f g t ) d t ,
ε ( t ) = 1 2 E ( t ) exp [ i ( ω 0 t + ϕ C E ) ] + c . c . ,
A f ( 2 ) = A a b s ( 2 ) + A R a m ( 2 ) ,
A a b s ( 2 ) = exp ( i 2 ϕ C E ) μ a b s 2 0 E ˜ ( ω ) E ˜ ( Ω f g ω ) d ω A R a m ( 2 ) = μ R a m , 1 2 Ω f g E ˜ ( ω ) E ˜ ( ω Ω f g ) d ω + μ R a m , 2 2 Ω f g E ˜ ( ω Ω f g ) E ˜ ( ω ) d ω ,
A f ( 3 ) ε 3 ( t ) exp ( i Ω f g t ) d t .
A f ( 3 ) = A a b s ( 3 ) + A R a m ( 3 ) ,
A a b s ( 3 ) = exp ( i 3 ϕ C E ) μ a b s 3 0 0 E ˜ ( ω 1 ) E ˜ ( ω 2 ) E ˜ ( Ω f g ω 1 ω 2 ) d ω 1 d ω 2
A R a m ( 3 ) = exp ( i ϕ C E ) [ μ R a m , 1 3 0 0 E ˜ ( ω 1 ) E ˜ ( ω 2 ) E ˜ ( ω 1 + ω 2 Ω f g ) d ω 1 d ω 2 + μ R a m , 2 3 0 0 E ˜ ( ω 1 + ω 2 Ω f g ) E ˜ ( ω 1 ) E ˜ ( ω 2 ) d ω 1 d ω 2 + μ R a m , 3 3 0 0 E ˜ ( ω 1 ) E ˜ ( ω 1 + ω 2 Ω f g ) E ˜ ( ω 2 ) d ω 1 d ω 2 ] ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.