Skip to main content

Main menu

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
    • Front Matter Portal
    • Journal Club
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
  • Submit
  • About
    • Editorial Board
    • PNAS Staff
    • FAQ
    • Accessibility Statement
    • Rights and Permissions
    • Site Map
  • Contact
  • Journal Club
  • Subscribe
    • Subscription Rates
    • Subscriptions FAQ
    • Open Access
    • Recommend PNAS to Your Librarian

User menu

  • Log in
  • My Cart

Search

  • Advanced search
Home
Home
  • Log in
  • My Cart

Advanced Search

  • Home
  • Articles
    • Current
    • Special Feature Articles - Most Recent
    • Special Features
    • Colloquia
    • Collected Articles
    • PNAS Classics
    • List of Issues
  • Front Matter
    • Front Matter Portal
    • Journal Club
  • News
    • For the Press
    • This Week In PNAS
    • PNAS in the News
  • Podcasts
  • Authors
    • Information for Authors
    • Editorial and Journal Policies
    • Submission Procedures
    • Fees and Licenses
  • Submit
Research Article

Pendular alignment and strong chemical binding are induced in helium dimer molecules by intense laser fields

Qi Wei, Sabre Kais, Tomokazu Yasuike, and Dudley Herschbach
  1. aState Key Laboratory of Precision Spectroscopy, East China Normal University, 200062 Shanghai, China;
  2. bDepartment of Chemistry, Purdue University, West Lafayette, IN 47907;
  3. cBirck Nanotechnology Center, Purdue University, West Lafayette, IN 47907;
  4. dDepartment of Liberal Arts, The Open University of Japan, 261-8586 Chiba, Japan;
  5. eElements Strategy Initiative for Catalysts and Batteries, Kyoto University, 615-8520 Kyoto, Japan;
  6. fDepartment of Physics, Texas A&M University, College Station, TX 77843;
  7. gDepartment of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138

See allHide authors and affiliations

PNAS September 25, 2018 115 (39) E9058-E9066; first published September 7, 2018; https://doi.org/10.1073/pnas.1810102115
Qi Wei
aState Key Laboratory of Precision Spectroscopy, East China Normal University, 200062 Shanghai, China;
bDepartment of Chemistry, Purdue University, West Lafayette, IN 47907;
cBirck Nanotechnology Center, Purdue University, West Lafayette, IN 47907;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: qwei@admin.ecnu.edu.cn dherschbach@gmail.com
Sabre Kais
bDepartment of Chemistry, Purdue University, West Lafayette, IN 47907;
cBirck Nanotechnology Center, Purdue University, West Lafayette, IN 47907;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Tomokazu Yasuike
dDepartment of Liberal Arts, The Open University of Japan, 261-8586 Chiba, Japan;
eElements Strategy Initiative for Catalysts and Batteries, Kyoto University, 615-8520 Kyoto, Japan;
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
Dudley Herschbach
fDepartment of Physics, Texas A&M University, College Station, TX 77843;
gDepartment of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138
  • Find this author on Google Scholar
  • Find this author on PubMed
  • Search for this author on this site
  • For correspondence: qwei@admin.ecnu.edu.cn dherschbach@gmail.com
  1. Contributed by Dudley Herschbach, July 29, 2018 (sent for review June 20, 2018; reviewed by Nimrod Moiseyev and Mitchio Okumura)

  • Article
  • Figures & SI
  • Info & Metrics
  • PDF
Loading

Significance

Intense electric fields, provided by pulsed lasers, can profoundly alter the electronic structure of atoms and molecules. For the helium dimer, we carry out a theoretical study of laser interactions in two realms: (I) fields not strong enough to dislodge electrons, but interact with the anisotropic polarizability to induce spatial alignment of the molecular axis; and (II) superintense, high-frequency lasers that impel electrons to undergo quiver oscillations that interact with the intrinsic Coulomb forces and induce an extremely strong chemical bond. By including in II an excited electronic state, we bring out features amenable to experimental observation that has been lacking.

Abstract

Intense pulsed-laser fields have provided means to both induce spatial alignment of molecules and enhance strength of chemical bonds. The duration of the laser field typically ranges from hundreds of picoseconds to a few femtoseconds. Accordingly, the induced “laser-dressed” properties can be adiabatic, existing only during the pulse, or nonadiabatic, persisting into the subsequent field-free domain. We exemplify these aspects by treating the helium dimer, in its ground (X1Σg+) and first excited (A1Σu+) electronic states. The ground-state dimer when field-free is barely bound, so very responsive to electric fields. We examine two laser realms, designated (I) “intrusive” and (II) “impelling.” I employs intense nonresonant laser fields, not strong enough to dislodge electrons, yet interact with the dimer polarizability to induce binding and pendular states in which the dimer axis librates about the electric field direction. II employs superintense high-frequency fields that impel the electrons to undergo quiver oscillations, which interact with the intrinsic Coulomb forces to form an effective binding potential. The dimer bond then becomes much stronger. For I, we map laser-induced pendular alignment within the X state, which is absent for the field-free dimer. For II, we evaluate vibronic transitions from the X to A states, governed by the amplitude of the quiver oscillations.

  • laser-induced properties
  • pendular alignment
  • quiver oscillations
  • chemical bonding
  • Kramers–Henneberger approximation

The 4He2 dimer, in its ground electronic state and field-free, is extremely fragile. Although its potential energy curve has appreciable well depth (Fig. 1A), the vibrational zero-point energy is nearly equal. Hence, the sole bound vibrational state lies only slightly below the separated atoms asymptote. The vibrational wavefunction thus extends far beyond the classical turning point, making the dimer extremely bloated. Its average internuclear distance is about 20× longer than that at the well minimum. With 80% probability, the two nuclei of the dimer reside outside the classically forbidden region. Determining accurately these properties has been a 25-y odyssey that led to exquisite advances in both theory (1⇓–3) and experiments (4⇓⇓⇓–8). A kindred subfield has developed, which treats “long-range” diatomic molecules, weakly bound in vibrational states that lie not far below the dissociation asymptote (9). Such states qualify as “quantum halos” as the vibrational wavefunction tunnels far into classically forbidden regions (10). Recently, the He2 ground-state halo wavefunction has been imaged with remarkable accuracy by extraordinary experiments using a Coulomb explosion technique (8).

Fig. 1.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 1.

Ground electronic state X1Σg+ of 4He2 dimer. (A) Field-free dimer potential energy curve (1–3). Well depth V(Rm)=3.5×10−5 a.u. (11 K) at Rm=5.61 a.u. Sole vibrational energy level (dashed blue) E0=−4.1×10−9 a.u. (−1.3 mK) is only slightly below asymptote for separated atoms. Probability distribution of internuclear distance (black, arbitrary units), R2|Ψvib(R)|2, extends beyond 100 a.u. Expectation value for internuclear distance 〈R〉 = 104 a.u. (B) Laser field (I = 0.005 a.u.)-induced alteration of potential energy curve (compared with free-field, dashed red), with downward shift of sole pendular–vibrational level (|000〉, dashed blue). (C) For stronger laser field (I = 0.01 a.u.). In B and C, the radial curves shown pertain to θ = 0°, whereas the penvib levels incorporate ranges of both R and θ; see Fig. 3 and Table 2.

Further field-free micro- and macroscopic properties of the ground X state of He2 have been studied (11). Also, field-free spectra have been obtained for several excited electronic states, particularly the A state (12). Previous theoretical work, applying superintense pulsed-laser fields (realm II) to ground-state He2 has predicted formation of a strong chemical bond (13⇓–15). However, the induced bond exists just during the pulse duration and experimental confirmation is as yet lacking. Here we treat theoretical aspects that bring out features amenable to experimental observation. These include treating less intense laser fields (realm I), which can provide pendular alignment and spectra (16⇓⇓⇓–20), and treating the A state in addition to the X state (in realm II) to obtain vibronic spectra.

Unless explicitly otherwise, we use atomic units: 1 a.u. for distance is a bohr unit (0.0529 nm); for time 2.42×10−17 s; for mass 9.11×10−25 kg; for energy a hartree unit (27.2 eV; equivalent to 3.16×105 kelvin); for laser intensity 3.51×1016 Wcm−2; and for laser frequency 6.58×1015 Hz, corresponding to wavelength of 45.5 nm.

Realm I: Laser Interactions with Molecular Polarizability

In this realm, the external laser field is less strong than the internal Coulombic forces that govern the electronic structure within a molecule. Laser fields then perturb the electronic structure of a typical molecule chiefly via its polarizability (16⇓⇓⇓⇓–21). For a nonpolar diatomic molecule with polarizability components α|| and α⊥ parallel and perpendicular to the molecular axis, subjected to plane-wave radiation of nonresonant frequency ω with electric field strength, ε = ε0cosωt, the interaction potential isVα(R,θ)=−1/2ε2g(t)[Δαcos2θ+α⊥].[1]Here g(t) is the time profile of the laser pulse with peak intensity I=ε2. For a Gaussian profile, g(t)=exp[−4ln(2)t2/τ2], the full width at half maximum τ is termed the pulse duration. For nonresonant frequencies much greater than the reciprocal of the laser pulse duration, ω≫1/τ, averaging over the pulse period converts ε2 to ε02/2. The dependence on the angle θ between the molecular axis and the electric field direction is governed by the anisotropy Δα=α||−α⊥ of the polarizability. The term Δαcos2θ introduces an equatorial barrier that quenches rotation but allows the molecular axis to undergo pendular libration about the electric field direction. The internuclear distance R enters implicitly as a parameter in the polarizability components. Also, we note that spectra governed by molecular polarizability have the same selection rules as the Raman effect (22); for a diatomic rotor, ΔJ=±2,0.

Before outlining calculations, we exhibit results for the ground-state helium dimer, resulting from Eq. 1, the interaction potential. Fig. 1 contrasts laser-induced changes (Fig. 1 B and C) in radial vibrational potential energy curves with the field-free case (Fig. 1A). The well depths are substantially deepened, along with consequent lowering of the sole bound level, which is properly designated a pendular–vibrational level (dashed blue). That lowering is accompanied by retreating of the probability distributions of the bond length. The radial potential curves and “penvib” levels shown incorporate θ, the alignment angle. Fig. 2A displays the dependence of the polarizability components on the internuclear distance. Fig. 2B exhibits the angular barrier imposed by the anisotropy of the polarizabilty, via Δαcos2θ. Fig. 3 presents 2D potential energy surfaces and wavefunctions to bring out the joint dependence of angular alignment along with the internuclear distance. Fig. 4 shows the dependence on the laser intensity (in Fig. 4A) of the penvib levels and (in Fig. 4B) of 〈cos2θ〉 and the pendular alignment range, Δθ, which narrows as the laser intensity is increased.

Fig. 2.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 2.

Quantities involved in interaction potential, Eq. 1, for ground-state helium dimer. (A) Dependence of polarizability components on internuclear distance. (Inset) Anisotropy is shown. (B) Laser-induced angular alignment of the dimer axis (schematic sketch). Potential minima in the polar regions, near θ = 0° and 180°, are separated by an equatorial barrier that quenches end-for-end rotation. Location of lowest penvib level is indicated by |000〉 (dashed blue).

Fig. 3.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 3.

Two-dimensional plots exhibiting radial and angular dependence, corresponding to Fig. 1. (A–C) Potential energy surfaces with transparent planes (green) that depict location of the lowest penvib quantum level. (A′–C′) Probability distributions, R2|Ψ(R, θ)|2, square of wavefunctions weighed by the radial Jacobian factor.

Fig. 4.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 4.

Variation with laser intensity of properties of ground-state helium dimer. (A) The laser-induced bound penvib levels EvJM = |000〉 and |021〉 (blue) include the contribution from the centrifugal term, 〈BJ2〉, in Eq. 2. The dotted curve shows where the |000〉 level would be if the centrifugal term were omitted. Also shown is the potential depth V(Rm). Numbers close to points indicate the averaged internuclear distance 〈R〉; compare Fig. 1 B and C and Table 2. (B) Expectation values 〈cos2θ〉 for the penvib levels (blue). Also indicated (±Δθ, black points) are angular amplitudes of the pendular librations, Δθ=arccos⁡〈cos2θ〉1/2.

Incorporating the interaction potential requires treating the full Hamiltonian,H=Hvib+Hrot+Vα(R,θ),[2]wherein the vibrational and rotational terms have the familiar field-free form:Hvib=−1/2 μd2/dR2+V0(R);Hrot=BJ2,[3]with μ (=mHe/2=3,675 a.u.) the reduced mass, V0(R) the field-free radial potential, and B=1/(2 μR2) the rotational constant. To obtain the penvib states from Eq. 2, we applied a stepwise iterative method. Step 1 sets cos2θ and J2 as zero, temporarily. Then R is the only variable, so a radial wavefunction ΨR is readily obtained by conventional means. In step 2, a given R is taken as constant with θ and J2 variable, and a pendular wavefunction Ψθ is obtained. The process is repeated for many values of R. In step 3, the product ΨRΨθ is used to calculate the expectation values 〈cos2θ〉 and 〈J2〉 and the energy levels. These are inserted into the full Hamiltonian and the whole procedure is iterated until the quantities obtained in step 3 converge; typically seven or eight iterations yield six-digit agreement.

The laser pulse duration has a key role (19⇓–21). Fig. 5A displays the variation with laser intensity of the helium dimer rotational period π/B along with contrasting shorter and longer pulse durations. Also shown is the constraint imposed by the ionization half-life T1/2 as estimated from atomic helium data (23). For example, for I ∼ 0.01 a.u., π/B ∼ 25 ps and T1/2 ∼ 10 ps. The designated pulse durations π/5B and 5π/B approach limits: nonadiabatic and adiabatic.

Fig. 5.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 5.

Variation with laser intensity of pulse options. (A) Helium dimer rotational period, π/B (blue curve), along with dashed curves that indicate shorter and longer pulse durations: Tpd=π/5B and 5π/B, respectively for nonadiabatic and adiabatic options. Also shown is the ionization half-life (red curve) of atomic helium, T1/2=ln⁡2/Γ, derived from single-electron ionization rates, for laser fields of wavelength 780 nm, obtained from accurate theoretical data of ref. 23. (B) Estimated fraction (%) of dimer molecules surviving ionization after undergoing laser pulses: exp(−0.6931Tpd/T1/2).

If the pulse is much shorter than π/B, the molecule is left in a coherent superposition of rotational eigenstates. Such an impulsive pulse can yield nonadiabatic behavior: The molecule retains the imparted pendular alignment and angular momentum after the laser pulse has passed. For He2, due to its extremely weak bond, the impulsive pulse can result in dissociation from “shaking” or “kicking” imposed by the centrifugal angular momentum (24⇓–26). This situation has been recently experimentally demonstrated (27). Such dissociation can occur whenever vibrational motion is unperturbed or slightly so, and the retained centrifugal energy exceeds the gap between the highest field-free vibrational level and the dissociation asymptote.

If the laser pulse is much longer than π/B, the molecule is adiabatically guided into pendular states but emerges from the pulse unchanged from the initial field-free state. Hence, the induced penvib states need to be observed midway in the pulse via a pump–probe technique. However, a hybrid technique is available (21, 28). The laser field can be adiabatically turned on, producing the penvib states, then suddenly turned off. The adiabatically prepared states can then dephase and rephase to form the same revival output that would be obtained from an impulsive pulse.

As shown in Fig. 5B, with increasing laser intensity depletion by ionization sets in much more drastically for the adiabatic and hybrid modes than for the impulsive mode. Up to I = 0.005 a.u., the dimer molecules that survive ionization are estimated to exceed 90% for all three pulse modes. But, by I = 0.010 a.u. the survival has fallen below 10−3% for the adiabatic and hybrid modes, yet holds up to ∼60% for the impulsive mode. Furthermore, at I = 0.015 a.u. the impulsive survival remains at ∼15%. Since ionized molecules can be readily deflected away, the surviving molecules can be readily recorded down to 1%, so the practical border for realm I occurs near I = 0.008 a.u. for the adiabatic mode and beyond I = 0.020 a.u. for the impulsive mode.

The laser-induced properties of He2 predicted in Fig. 4 thus are accessible for the impulsive mode. The |000〉 penvib level is lowered enough by the laser interaction to offset the extra centrifugal energy, thus avoiding dissociation. Likewise, the |021〉 level also becomes bound when the laser intensity climbs above 0.013 a.u. (The |020〉 and |022〉 levels remain unbound until the laser intensity exceeds 0.030 a.u.) The transition between |000〉 and |021〉 levels offers a compelling test. It occurs in the millimeter wave region, ranging from 6–11 cm−1 (180–330 GHz) for laser intensity from 4–7×1014 W/cm2 (0.01–0.02 a.u.). Suitable probe lasers for this region are available (29). Lower limit of probe pulse duration ranges from 3 to 5 ps. The alignment of the dimer axis, recorded by 〈cos2θ〉, provides another test, accessible by particle-imaging techniques (30).

Realm II: Laser Disruption of Electronic Structure

In this realm the laser field is superintense, comparable to the intrinsic Coulomb binding forces, and well above the threshold for ionization. However, theory and computer simulations predict that in this regime the ionization probability decreases as the laser intensity increases (31⇓–33). This apparent paradox occurs because the laser-dressed molecule acquires an effective binding potential for the electrons by interaction of the rapidly oscillating laser field with the Coulombic potential. The stabilization against ionization even extends to multiply charged anions of hydrogen (34) and anions of other atoms (35⇓–37), as well as positronium (38) and some simple diatomic molecules (13⇓–15, 39). Even more striking is a theoretical demonstration, using a carefully devised laser dressing that counteracts Coulombic repulsion, to form a metastable HD2+ molecule (40). That emulates a juggler, balancing a stick on the tip of a finger. Kindred stabilization effects appear in the inverse pendulum of Kapitza (41), in the Paul mass filter (42), and Hau (43) guided matter waves.

The stabilization regime depends on the laser frequency, as well as the laser intensity. When the quiver oscillations of electrons driven by the laser field become dominant, a juggler-like procedure termed high-frequency Floquet theory (HFFT) has provided a good approximation for time averaging (31⇓–33). As a small molecule, such as He2, is miniscule compared with the wavelength of the laser light, each of its electrons is subject to the same laser field. All then undergo synchronously quiver oscillations,a(t)=α0⁡cos(ωt)e^,[4]along the electric field vector (for linearly polarized light e^ is a unit vector orthogonal to the propagation direction). The maximum quiver amplitude is α0=E0/ω2, with ω the frequency and E0=I1/2 the field amplitude. The HFFT procedure is simplified by adopting a reference frame attached to the quivering electrons, designated the Kramers–Henneberger (KH) frame (44, 45). It is translated by a(t) with respect to the laboratory frame. Hence, in the KH frame the electrons all remain fixed, while instead the nuclei quiver along the a(t) trajectory. Thereby the Coulombic attraction between any electron and a nucleus with charge Z takes the form −Z/|ri+α(t)|. The electrons then feel a time-averaged effective attractive potential, the “dressed” potential, given byVKH(ri,α0)=−ω2π∫02π/ωZdt|ri+α(t)|,[5]where the time average extends over one period of the laser field. There are higher-order frequency-dependent corrections to the dressed potential, which are proved to be small and can be neglected at high frequencies (36).

The HFFT version of the time-independent electronic Schrӧdinger equation, in accord with the Born–Oppenheimer approximation for a homonuclear diatomic molecule, has the familiar form for a field-free molecule, except for replacing the electron-nucleus terms with the dressed KH potential terms:∑i=1N[12pi2+VKH(ri−R2,α0)+VKH(ri+R2,α0)+∑j=1i−11|ri−rj|+Z2R]Φ=ϵ(N)(α0,R)Φ.[6]The laser intensity and frequency thus appear only in the KH terms and enter only via the quiver amplitude. The energy eigenvalues ϵ(N)(α0,R) and wavefunctions Φ are functions of the number N of electrons, the quiver amplitude α0, and the internuclear vector R, both its magnitude and its angle θ from the electric field direction. Computational details in evaluating the KH integrals of Eq. 6 are described in previous papers (13⇓–15). We inserted the integrals obtained in ref. 13 into the standard GAMESS program package (46) and employed conventional Hartree–Fock orbitals (restricted Hartree–Fock, RHF, rather than unrestricted HF). For the range of quiver amplitudes that we treated, α0≤2, comparisons with accurate numerical calculations verified that this procedure was adequate for the ground electronic state, X1Σg+. In treating the first excited electronic state, A1Σu+, we augmented the RHF orbitals with a single configuration interaction (CIS) approximation (47). The CIS method is considered to provide a well-balanced description for one-electron exited states compared with an HF ground state.

The pair of KH terms in Eq. 6 can be viewed (48) as the electrostatic potentials generated by two lines of positive charge, depicted in Fig. 6A, each of length 2α0, parallel to the electric field vector and centered on the nuclei at ±R/2. The electrons hence are attracted to the “smeared-out” lines of the nuclear charges and are most attracted to the quiver endpoints at ±α0. Consequently, superlaser intensities induce the electron distribution to exhibit dichotomy (49). The dressed Coulombic potentials cluster electrons near the quiver endpoints, as if they were a pair of virtual nuclei separated by 2α0. For He2, the dichotomy is feeble for α0≤0.2 and modest for quiver amplitudes below α0≤2. At larger amplitudes, the lobes become more widely spaced and foster stabilization by decreasing the ionization rate.

Fig. 6.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 6.

(A) “Lines of charge” generating the effective potential acting on each electron in the VKH terms in Eq. 6. These line segments of length 2α0 are parallel to the laser polarization along the z axis (toward unit vector e of Eq. 4) and centered on the nuclei at ±R/2. The internuclear vector R is directed at angle θ from the z axis, here drawn with the azimuth ϕ = 0 about the z axis (eigenproperties do not depend on that uniform angle). (B) Variation with θ of potential well depth: V(α0,Rm,θ)=ϵ(4)(α0,Rm,θ)−ϵ(4)(α0,∞,θ), from eigenvalues of Eq. 6, for X (red) and A (blue) states and for different quiver amplitudes α0=0.2,0.5,1.0,1.5 and 2.0 a.u., respectively. Dots indicate, for the lowest penvib levels, the pendular range of the dimer axis at the radial minimum, Rm.

The eigenvalues of Eq. 6 provide the dimer potential energy surfaces: V(α0,R,θ). A striking feature, seen in Fig. 6B, is that for both the X and A states the pendular angular range of the dimer axis is constrained within less than ±10° when α0 = 0.5 and less than ±5° when α0≤1, for the lowest penvib levels and R=Rm, the potential minimum. Hence, in exploring the KH regime, we usually just set θ = 0. The potential energy surfaces shown in Fig. 7 have set θ = 0 in the upper panels but not in the lower panels. The most dramatic aspect is that with α0 ∼ 1–2 a.u., chemical bonding is drastically enhanced. That had been found previously for the X state (13⇓–15). The A state has substantial bonding when field-free, but we find it also is much enhanced. The results for both X and A states were computed numerically, but the potential curves V(R,θ=0) fit tolerably well to the Morse potential,V(R)=De(e−2β(R−Re)−2e−β(R−Re)).[7]Table 1 lists the three fitted parameters: β; D=V(Rm), the well depth; and Rm, the location of its minimum. For α0=2 a.u., taking the bond strength as about equal to the well depth expressed in a familiar chemical unit, the bonds DX=1,300 and DA=1,600 kJ/mol are extremely strong. Also notable is the large lowering of the asymptote for separated atoms for the A state compared with the X state.

Fig. 7.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 7.

KH 4He2 dimer potential energy surfaces for both electronic ground-state X1Σg+ and excited-state A1Σu+, for quiver amplitudes α0= 0, 0.5, 1, 2. (A–D, Upper) shows radial R dependence with optimal alignment (θ = 0) compared with field-free (dashed red); note that the zero for the ordinate energy scale is the asymptote for the ground-state separated atoms. (A′–D′, Lower) shows 2D plots of R, θ dependence (Table 1).

View this table:
  • View inline
  • View popup
Table 1.

Morse fit parameters for KH potential energy curves

In the eigenenergies obtained from Eq. 6, the laser intensity and frequency enter only via the quiver amplitude, α0=I/ω2, but the HFFT criteria (31) impose upper and lower bounds on the pump laser frequency:137/α0≫ω≫|ϵ(N)(α0,R)|/N.[8]The upper bound is required to ensure that both the dipole approximation holds (radiation wavelength large compared with 2α0) and a nonrelativistic treatment suffices (maximum quiver speed much lower than the speed of light). The lower bound requires that the field must oscillate much faster than electron motion within the molecule. This is specified by an average excitation energy in the field (48), usually estimated by the energy eigenvalue that appears on the right-hand side of Eq. 6 divided by the number of electrons in the molecule (N = 4). As displayed in Fig. 8, the lower bound evaluated at Rm is similar for the X and A states. For the laser intensity, estimates have been made for the X state of ω = 10 at α0 = 1.0 (50) and ω = 4 at α0 = 2.5 (15), but for the A state no estimate is available. Here we presume both states are roughly similar and simply exhibit ω = 5 and 10 (dashed red lines), which are well above the lower bound and below the upper bound, provided that α0 does not exceed 2 or 3. The corresponding laser intensity (dashed blue curves) increases strongly with increase of the quiver amplitude and even more strongly with the frequency. Consequently, e.g., if α0 = 1 and ω = 10, the requisite laser intensity should be 104a.u.=3.5×1020 W/cm2 with wavelength λ=45.5/ω=4.5 nm. Such properties can be attained from X-ray free-electron lasers, now accessible in several large facilities (51). Also, promising results obtaining coherent X-rays from high-harmonics generation have emerged from table-top experiments (52).

Fig. 8.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 8.

Criteria of Eq. 8 governing quiver amplitude and specifying lower and upper bounds (black curves) for laser frequency, with trial values ω = 10 and 5 (red dashed lines) and corresponding laser intensity (blue curves).

Illustrative Franck–Condon vibronic transitions between the X and A states are shown in Fig. 9, for α0 = 0.5 and α0 = 1. As seen in Fig. 7 and Table 1, for α0 = 0.5 the radial potential of the X state near R ∼ 2 is steeply repulsive below a deep well in the A state. Hence, there occurs an internal photodissociation. The emitted transition from vA = 0 reflects from the steep X wall, dissociates the dimer, and travels away as a continuous wave. The atoms fly apart, sharing the kinetic energy, which arises from the height of the impact point on the X potential above the exit asymptote. In contrast, near R ∼ 4 the X state has a modest well, so either an absorption or emission transition connects the vX = 0 level to the turning point of vA = 17, among other A vibrational levels. For α0 = 1, the X state has acquired a deep well near R ∼ 2 that aligns with the A state well, so both states offer transitions between ladders of discrete vibrational levels.

Fig. 9.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 9.

Franck–Condon transitions between KH potential curves of ground-state X1Σg+ and excited-state A1Σu+. (A and A′) For α0=0.5. From lowest vibrational level of the A state, vA = 0, transition reflects from the steep repulsive region of the X state, dissociates the dimer, and becomes a continuous traveling wave carrying off kinetic energy. Sample transition from higher level, vA = 17, connects turning point to the lowest vibrational level of the X state. Transition energies (units 103 cm−1): continuum peaks at 70, reflecting shape of vA=0, while discrete lines between 116 and 132 reflect higher vA levels with turning points that align with vX=0 (B and B′). For α0=1.0, both X and A potential curves acquire much deeper wells with similar Rm, show only 0 → 0 transition, shortest but most intense.

Our prime interest in the transitions between X and A is considering experiments that can confirm the KH Floquet approach. Much depends on the photoionization lifetime. Theoretical calculations (50) for the X state, when α0≈1 and ω=10 at fixed R=Rm, found the lifetime reached 1 ps at maximum but is very sensitive to R and decreases drastically away from Rm. The lifetime drops to a few femtoseconds for vibrational states. Also, except for the vX = 0 level, transitions (mid-IR region) between the adjacent vibrational state levels were overcome by broadening. In the much larger electronic transitions between X and A (UV region) the lifetimes allow distinct vibrational levels. As seen in Fig. 7, with increasing α0>0.7, the energy gap between the X and A potential surfaces decreases markedly. That change should provide clear evidence for KH dressing, available from either an absorption or emission discrete spectrum.

A complementary aspect, with decreasing α0<0.7, pursues continuum emission spectra between the A and X potential surfaces, as depicted in Fig. 10A. The field-free spectrum, for α0=0, is centered at 130×103 cm−1. It was observed nearly a century ago among UV continuum bands and soon identified as primarily from the A1Σu+ state and named for Hopfield (53). For α0 = 0.5, the continuum band becomes centered at 70×103 cm−1. Such a large shift in the spectrum offers a direct experimental test of the KH process. Moreover, the accompanying dissociation of the dimer liberates kinetic energy, as displayed in Fig. 10B. That invites another experimental test, detecting trajectories of the emerging helium atoms by means of an imaging technique widely used for photodissociation and reactive molecular collisions (30). At first blush, the predicted distribution of kinetic energy is not much different between α0 = 0 and α0 = 0.5, so the peak exit velocity of the He atoms should be similar. However, in our case of “internal photodissociation,” the field-free A state delivers spontaneous emission, with a leisurely lifetime of 0.55 ns (53). Instead, the KH regime is clearly distinct since it uses superintense lasers that deliver stimulated emission with a hasty lifetime below femtoseconds.

Fig. 10.
  • Download figure
  • Open in new tab
  • Download powerpoint
Fig. 10.

Internal photodissociation of He2 dimer. (A) Franck–Condon (FC) continuum transitions from KH potentials for α0=0,0.2,0.5 (blue, green, red; Fig. 9A) compared with field-free spectrum (black profile). Notation FC2 designates sum of squares of 〈vA|vX〉. (B) Corresponding kinetic energy carried off as the pair of helium atoms fly apart. The predicted exit velocity of each He atom at the peak distribution for α0=0,0.2,0.5 are similar: 12.5, 13.6, 11.9 km/s, respectively.

Discussion and Outlook

In realm I, we considered the interaction of nonresonant laser fields with the polarizability of the ground electronic state (X1Σg+) of the helium dimer (Figs. 1–4 and Table 2). The field-free dimer has only one barely bound vibrational level; in kelvin units, it lies only about 10−3 K below the He + He dissociation asymptote. The laser field enhances the molecular binding, but the sole level remains in the van der Waals range. For example, applying laser intensity of 3×1014W/cm2 merely lowers the location of the sole level to about 10 K. However, such laser intensity interacting with the anisotropy of the polarizability constrains the dimer axis to pendular librating motion within ±27°, a rather narrow range (Fig. 4). Accordingly, the sole bound level is appropriately designated penvib. Our chief concerns dealt with three issues mutually involved: (i) the key role of the laser pulse duration (Fig. 5); (ii) intrusion of ionization; and (iii) the centrifugal energy associated with the sole penvib level. The outcome, not anticipated, emerged that the laser interaction lowered the penvib level enough to offset the centrifugal input, hence avoiding dissociation.

View this table:
  • View inline
  • View popup
Table 2.

Laser-induced parameters for ground state of 4He2 dimer

In realm II, abundant theoretical work has adopted the KH approach since it emerged 50 y ago (45). However, experimental confirmation for KH effects has been meager, limited mostly to Rydberg atoms until the recent decade (54, 55). As yet, experimental affirmation is still lacking for molecules, particularly the helium dimer. Previous theoretical pursuits of KH dressing have all dealt with the ground electronic state, assessing bond energies, molecular orbitals, and vibrational and rotational levels (13⇓–15). We have introduced the excited electronic state (A1Σu+) to provide more accessible experimental tests with a wide range of discrete and continuum spectra (Figs. 6–10 and Table 1). In particular, the transitions between the X and A states occur in times much shorter than molecular rotation or pendular motion, so the transition dipole moment aligns with the electric vector of the light (56). Accordingly, in the continuum spectra the laser polarization supplies a distinctive angular distribution of the departing atoms.

Realms I and II share the same experimental techniques in producing a molecular beam of the helium dimer and detecting it. In an elegant procedure, a pure helium dimer beam cooled to 8 K has been obtained by matter wave diffraction (8). Also elegant is detecting the beam by ionizing singly the atoms, so the two positively charged atoms repel, resulting in a Coulomb explosion. Recording the momentum vectors of the ions provides velocity and directional data. The process was recently exemplified for the dimer field-free X state by imaging the wavefunction and binding (8). It has been used also to study nonadiabatic alignment of the dimer (27). The imaging process noted in our discussion of A–X internal photodissociation does not require a Coulomb explosion (30). These particle-imaging techniques can now cope with the continuum spectra while well-established pump–probe techniques deal with the discrete spectra in elucidating the KH regime.

Methods

In realm I, a self-consistent stepwise iterative method was used to obtain the penvib states from Eq. 2. Radial and pendular parts of the equation were solved separately. In realm II, RHF and CIS methods were used to obtain potential surfaces of X1Σg+ and A1Σu+ states from Eq. 6. The method used for penvib states in realm II is the same as that in realm I.

Acknowledgments

Q.W. thanks Konstantin Dorfman for enlightening discussion of lasers. We are grateful for support from National Natural Science Foundation of China (Grant 11674098) and Institute for Quantum Science and Engineering, Texas A&M University.

Footnotes

  • ↵1To whom correspondence may be addressed. Email: qwei{at}admin.ecnu.edu.cn or dherschbach{at}gmail.com.
  • Author contributions: Q.W., S.K., T.Y., and D.H. designed research; Q.W., T.Y., and D.H. performed research; Q.W., S.K., and D.H. analyzed data; and Q.W. and D.H. wrote the paper.

  • Reviewers: N.M., Technion Israel Institute of Technology; and M.O., California Institute of Technology.

  • The authors declare no conflict of interest.

Published under the PNAS license.

References

  1. ↵
    1. Anderson JB,
    2. Trynor CA,
    3. Boghosian BM
    (1993) An exact quantum Monte Carlo calculation of the helium-helium intermolecular potential. J Chem Phys 99:345–351.
    OpenUrl
  2. ↵
    1. Tang KT,
    2. Toennies JP,
    3. Yiu CL
    (1995) Accurate analytical He-He van der Waals potential based on perturbation theory. Phys Rev Lett 74:1546–1549.
    OpenUrlCrossRefPubMed
  3. ↵
    1. Przybytek M, et al.
    (2010) Relativistic and quantum electrodynamics effects in the helium pair potential. Phys Rev Lett 104:183003.
    OpenUrlPubMed
  4. ↵
    1. Lou F,
    2. McGane GC,
    3. Kirn G,
    4. Giese CF,
    5. Gentry WR
    (1993) The weakest bond: Experimental observation of the helium dimer. J Chem Phys 98:3564–3567.
    OpenUrlCrossRef
  5. ↵
    1. Luo F,
    2. Giese CF,
    3. Gentry WR
    (1996) Direct measurement of the size of the helium dimer. J Chem Phys 104:1151–1154.
    OpenUrlCrossRef
  6. ↵
    1. Schollkopf W,
    2. Toennies JP
    (1996) The nondestructive detection of the helium dimer and trimer. J Chem Phys 104:1155–1158.
    OpenUrl
  7. ↵
    1. Grisenti RE, et al.
    (2000) Determination of the bond length and binding energy of the helium dimer by diffraction from a transmission grating. Phys Rev Lett 85:2284–2287.
    OpenUrlCrossRefPubMed
  8. ↵
    1. Zeller S, et al.
    (2016) Imaging the He2 quantum halo state using a free electron laser. Proc Natl Acad Sci USA 113:14651–14655.
    OpenUrlAbstract/FREE Full Text
  9. ↵
    1. Zemke WT,
    2. Byrd JN,
    3. Michels HH,
    4. Montgomery JA Jr,
    5. Stwalley WC
    (2010) Long range intermolecular interactions between the alkali diatomics Na(2), K(2), and NaK. J Chem Phys 132:244305.
    OpenUrlPubMed
  10. ↵
    1. Jensen AS,
    2. Rusagr K,
    3. Fedorov DV
    (2004) Structure and reactions of quantum halos. Rev Mod Phys 76:215–261.
    OpenUrlCrossRef
  11. ↵
    1. Komasa J
    (2001) Exponentially correlated Gaussian functions in variational calculations. Momentum space properties of the ground state helium dimer. J Chem Phys 115:158–165.
    OpenUrl
  12. ↵
    1. Focsa C,
    2. Bernath PF,
    3. Colin R
    (1998) The low-lying states of He2. J Mol Spectrosc 191:209–214.
    OpenUrlPubMed
  13. ↵
    1. Yasuike T,
    2. Someda K
    (2004) He-He chemical bonding in high-frequency intense laser fields. J Phys B At Mol Opt Phys 37:3149–3162.
    OpenUrl
  14. ↵
    1. Wei Q,
    2. Kais S,
    3. Herschbach D
    (2008) Dimensional scaling treatment of stability of simple diatomic molecules induced by superintense, high-frequency laser fields. J Chem Phys 129:214110.
    OpenUrlPubMed
  15. ↵
    1. Balanarayan P,
    2. Moiseyev N
    (2012) Strong chemical bond of stable He2 in strong linearly polarized laser fields. Phys Rev A 85:032516.
    OpenUrl
  16. ↵
    1. Friedrich B,
    2. Herschbach D
    (1995) Alignment and trapping of molecules in intense laser fields. Phys Rev Lett 74:4623–4626.
    OpenUrlCrossRefPubMed
  17. ↵
    1. Friedrich B,
    2. Herschbach D
    (1995) Polarization of molecules induced by intense nonresonant laser fields. J Phys Chem 99:15686–15693.
    OpenUrl
  18. ↵
    1. Seideman T
    (1995) Rotational excitation and molecular alignment in intense laser fields. J Chem Phys 103:7887–7896.
    OpenUrl
  19. ↵
    1. Ortigoso J,
    2. Rodriguez M,
    3. Gupta M,
    4. Friedrich B
    (1999) Time evolution of pendular states created by the interaction of molecular polarizability with a pulsed nonresonant laser field. J Chem Phys 110:3870–3875.
    OpenUrlCrossRef
  20. ↵
    1. Stapelfeldt H,
    2. Seideman T
    (2003) Colloquium: Aligning molecules with strong laser pulses. Rev Mod Phys 75:543–557.
    OpenUrlCrossRef
  21. ↵
    1. Torres R,
    2. de Nalda R,
    3. Marangos JP
    (2005) Dynamics of laser-induced molecular alignment in the impulsive and adiabatic regimes: A direct comparison. Phys Rev A 72:023420.
    OpenUrl
  22. ↵
    1. Condon EU
    (1932) Production of infrared spectra with electric fields. Phys Rev 41:759–762.
    OpenUrl
  23. ↵
    1. Parker JS,
    2. Meharg KJ,
    3. McKenna GA,
    4. Taylor KT
    (2007) Single-ionization of helium at Ti:Sapphire wavelengths: Rates and scaling laws. J Phys B At Mol Opt Phys 40:1729–1743.
    OpenUrl
  24. ↵
    1. Friedrich B,
    2. Gupta M,
    3. Herschbach D
    (1998) Probing weakly bound species with nonresonant light: Dissociation of He2 induced by rotational hybridization. Collect Czech Chem Commun 63:1089–1093.
    OpenUrl
  25. ↵
    1. Lemeshko M,
    2. Friedrich B
    (2009) Probing weakly bound molecules with nonresonant light. Phys Rev Lett 103:053003.
    OpenUrlPubMed
  26. ↵
    1. Lemeshko M,
    2. Friedrich B
    (2009) Rotational and rotationless states of weakly bound molecules. Phys Rev A 79:050501.
    OpenUrl
  27. ↵
    1. Kunitski M
    (2018) Rotating rotationless: Nonadiabatic alignment of the helium dimer (Goethe Universitat, Frankfurt Am Main), in press.
  28. ↵
    1. Yan ZC,
    2. Seideman T
    (1999) Photomanipulation of molecular modes: A time-dependent self-consistent-field approach. J Chem Phys 111:4113–4120.
    OpenUrl
  29. ↵
    1. Weber MJ
    (1999) Handbook of Laser Wavelengths (CRC Press, Boca Raton, FL).
  30. ↵
    1. Chandler DW,
    2. Houston PL,
    3. Parker DH
    (2017) Perspective: Advanced particle imaging. J Chem Phys 147:013601.
    OpenUrl
  31. ↵
    1. Gavirla M
    , ed (1992) Atoms in Intense Laser Fields (Academic, New York).
  32. ↵
    1. Eberly JH,
    2. Kulander KC
    (1993) Atomic stabilization by super-intense lasers. Science 262:1229–1233.
    OpenUrlAbstract/FREE Full Text
  33. ↵
    1. Gavila M
    (2002) Topical review: Atomic stabilization in superintense laser fields. J Phys B At Mol Opt Phys 35:R147–R193.
    OpenUrlCrossRef
  34. ↵
    1. van Duijn E,
    2. Gavrila M,
    3. Muller HG
    (1996) Multiply charged negative ions of hydrogen induced by superintense laser fields. Phys Rev Lett 77:3759–3762.
    OpenUrlPubMed
  35. ↵
    1. Wei Q,
    2. Kais S,
    3. Moiseyev N
    (2006) New stable multiply charged negative atomic ions in linearly polarized superintense laser fields. J Chem Phys 124:201108.
    OpenUrlPubMed
  36. ↵
    1. Wei Q,
    2. Kais S,
    3. Moiseyev N
    (2007) Frequency-dependent stabilization of He− by a superintense laser field. Phys Rev A 76:013407.
    OpenUrl
  37. ↵
    1. Wei Q,
    2. Kais S,
    3. Herschbach D
    (2007) Dimensional scaling treatment of stability of atomic anions induced by superintense, high-frequency laser fields. J Chem Phys 127:094301.
    OpenUrlPubMed
  38. ↵
    1. Wei Q,
    2. Herschbach D
    (2013) Positronium in superintense high-frequency laser fields. Mol Phys 111:1835–1843.
    OpenUrl
  39. ↵
    1. Balanarayan P,
    2. Moiseyev N
    (2013) Chemistry in high-frequency strong laser fields: The story of HeS molecule. Mol Phys 111:1814–1822.
    OpenUrl
  40. ↵
    1. Smirnova O,
    2. Spanner M,
    3. Ivanov M
    (2003) Molecule without electrons: Binding bare nuclei with strong laser fields. Phys Rev Lett 90:243001.
    OpenUrlCrossRefPubMed
  41. ↵
    1. Ter Haar D
    1. Kapitza PI
    (1965) in Collected Papers, ed Ter Haar D (Pergamon, Oxford), Vol 2.
  42. ↵
    1. Paul W
    (1990) Electromagnetic traps for charged and neutral particles. Rev Mod Phys 62:531–540.
    OpenUrlCrossRef
  43. ↵
    1. Hau LV,
    2. Burns MM,
    3. Golovchenko JA
    (1992) Bound states of guided matter waves: An atom and a charged wire. Phys Rev A 45:6468–6478.
    OpenUrlPubMed
  44. ↵
    1. Kramers HA
    (1956) Collected Scientific Papers (North-Holland, Amsterdam), p 866.
  45. ↵
    1. Henneberger WC
    (1968) Perturbation method for atoms in intense light beams. Phys Rev Lett 21:838–841.
    OpenUrlCrossRef
  46. ↵
    1. Schmidt MW, et al.
    (1993) General atomic and molecular electronic structure system. J Comput Chem 14:1347–1363.
    OpenUrlCrossRef
  47. ↵
    1. Foreman JB,
    2. Head-Gordon M,
    3. Pople JA,
    4. Frisch MJ
    (1992) A systematic molecular orbital theory for excited states. J Phys Chem 96:135–149.
    OpenUrlCrossRef
  48. ↵
    1. Shertzer J,
    2. Chandler A,
    3. Gavrila M
    (1994) H2+ in superintense laser fields: Alignment and spectral restructuring. Phys Rev Lett 73:2039–2042.
    OpenUrlPubMed
  49. ↵
    1. Nguyen NA,
    2. Nguyen-Dang T-T
    (2000) Molecular dichotomy within an intense high-frequency laser field. J Chem Phys 112:1229–1239.
    OpenUrl
  50. ↵
    1. Yasuike T,
    2. Someda K
    (2008) Lifetime of metastable helium molecule in intense laser fields. Phys Rev A 78:013403.
    OpenUrl
  51. ↵
    1. Pellegrini C,
    2. Marinelli A,
    3. Reiche S
    (2016) The physics of x-ray free-electron lasers. Rev Mod Phys 88:015006.
    OpenUrl
  52. ↵
    1. Gardner D, et al.
    (2017) Sub-wavelength coherent imaging of periodic samples using a 13.5 nm tabletop high harmonic light source. Nat Photonics 11:259–263.
    OpenUrl
  53. ↵
    1. Hill PC
    (1989) Ultraviolet continua of helium molecules. Phys Rev A Gen Phys 40:5006–5016.
    OpenUrl
  54. ↵
    1. Morales F,
    2. Richter M,
    3. Patchkovskii S,
    4. Smirnova O
    (2011) Imaging the Kramers-Henneberger atom. Proc Natl Acad Sci USA 108:16906–16911.
    OpenUrlAbstract/FREE Full Text
  55. ↵
    1. Wei Q,
    2. Wang P,
    3. Kais S,
    4. Herschbach D
    (2017) Pursuit of the Kramers-Henneberger atom. Chem Phys Lett 683:240–246.
    OpenUrl
  56. ↵
    1. Zare RN,
    2. Herschbach DR
    (1963) Doppler line shape of atomic fluorescence excited by molecular dissociation. Proc IEEE 51:173–182.
    OpenUrl
PreviousNext
Back to top
Article Alerts
Email Article

Thank you for your interest in spreading the word on PNAS.

NOTE: We only request your email address so that the person you are recommending the page to knows that you wanted them to see it, and that it is not junk mail. We do not capture any email address.

Enter multiple addresses on separate lines or separate them with commas.
Pendular alignment and strong chemical binding are induced in helium dimer molecules by intense laser fields
(Your Name) has sent you a message from PNAS
(Your Name) thought you would like to see the PNAS web site.
CAPTCHA
This question is for testing whether or not you are a human visitor and to prevent automated spam submissions.
Citation Tools
Pendular alignment and strong chemical binding are induced in helium dimer molecules by intense laser fields
Qi Wei, Sabre Kais, Tomokazu Yasuike, Dudley Herschbach
Proceedings of the National Academy of Sciences Sep 2018, 115 (39) E9058-E9066; DOI: 10.1073/pnas.1810102115

Citation Manager Formats

  • BibTeX
  • Bookends
  • EasyBib
  • EndNote (tagged)
  • EndNote 8 (xml)
  • Medlars
  • Mendeley
  • Papers
  • RefWorks Tagged
  • Ref Manager
  • RIS
  • Zotero
Request Permissions
Share
Pendular alignment and strong chemical binding are induced in helium dimer molecules by intense laser fields
Qi Wei, Sabre Kais, Tomokazu Yasuike, Dudley Herschbach
Proceedings of the National Academy of Sciences Sep 2018, 115 (39) E9058-E9066; DOI: 10.1073/pnas.1810102115
del.icio.us logo Digg logo Reddit logo Twitter logo CiteULike logo Facebook logo Google logo Mendeley logo
  • Tweet Widget
  • Facebook Like
  • Mendeley logo Mendeley

Article Classifications

  • Physical Sciences
  • Chemistry
Proceedings of the National Academy of Sciences: 115 (39)
Table of Contents

Submit

Sign up for Article Alerts

Jump to section

  • Article
    • Abstract
    • Realm I: Laser Interactions with Molecular Polarizability
    • Realm II: Laser Disruption of Electronic Structure
    • Discussion and Outlook
    • Methods
    • Acknowledgments
    • Footnotes
    • References
  • Figures & SI
  • Info & Metrics
  • PDF

You May Also be Interested in

Smoke emanates from Japan’s Fukushima nuclear power plant a few days after tsunami damage
Core Concept: Muography offers a new way to see inside a multitude of objects
Muons penetrate much further than X-rays, they do essentially zero damage, and they are provided for free by the cosmos.
Image credit: Science Source/Digital Globe.
Water from a faucet fills a glass.
News Feature: How “forever chemicals” might impair the immune system
Researchers are exploring whether these ubiquitous fluorinated molecules might worsen infections or hamper vaccine effectiveness.
Image credit: Shutterstock/Dmitry Naumov.
Venus flytrap captures a fly.
Journal Club: Venus flytrap mechanism could shed light on how plants sense touch
One protein seems to play a key role in touch sensitivity for flytraps and other meat-eating plants.
Image credit: Shutterstock/Kuttelvaserova Stuchelova.
Illustration of groups of people chatting
Exploring the length of human conversations
Adam Mastroianni and Daniel Gilbert explore why conversations almost never end when people want them to.
Listen
Past PodcastsSubscribe
Horse fossil
Mounted horseback riding in ancient China
A study uncovers early evidence of equestrianism in ancient China.
Image credit: Jian Ma.

Similar Articles

Site Logo
Powered by HighWire
  • Submit Manuscript
  • Twitter
  • Facebook
  • RSS Feeds
  • Email Alerts

Articles

  • Current Issue
  • Special Feature Articles – Most Recent
  • List of Issues

PNAS Portals

  • Anthropology
  • Chemistry
  • Classics
  • Front Matter
  • Physics
  • Sustainability Science
  • Teaching Resources

Information

  • Authors
  • Editorial Board
  • Reviewers
  • Subscribers
  • Librarians
  • Press
  • Cozzarelli Prize
  • Site Map
  • PNAS Updates
  • FAQs
  • Accessibility Statement
  • Rights & Permissions
  • About
  • Contact

Feedback    Privacy/Legal

Copyright © 2021 National Academy of Sciences. Online ISSN 1091-6490