National Academies Press: OpenBook
« Previous: 2 Design and Synthesis of Molecular Qubit Systems
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

3

Measurement and Control of Molecular Quantum Systems

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

As stated in Chapters 1 and 2, the degree of coherence in molecular systems is critical for chemical applications of QIS. Over the past two decades, there has been great progress not only in the discovery of new chemical designs and synthetic approaches but also in the development of new experimental measurements to probe the fundamental processes in molecules important for QIS applications. While there have been many experimental approaches in this regard, the bulk of the reported work in this area consists of magnetic and optical methods. The principle method to probe the spin dynamics and T1 and T2 relaxation process in organic and inorganic molecules has involved electron paramagnetic resonance (EPR) or other nuclear magnetic resonance (NMR) effects. Both time-resolved and continuous-wave (CW) EPR methods have had great impacts in the development of QIS chemical systems. A relatively new approach of using optical cycling in small molecules has also gained great attention for addressing and manipulating electron spin states. For optical measurements of coherence and QIS applications, the focus has been directed toward ultrafast, time-resolved, nonlinear optical spectroscopy and spectroscopy with quantum light (e.g., entangled photons). Two-dimensional (2D) time-resolved methods for investigations of organic and biological systems have been developed over the past two decades. In the case of quantum light applications, new methods for higher-yield entangled photon sources, nonlinear spectroscopy and microscopy with entangled photons, and quantum interferometry and the use of microcavities have been developed for chemical applications in QIS.

Chapter 3 provides an overview of the state of the art in the measurement of quantum phenomena using experimental chemical approaches and how quantum tools could be used to study chemical systems. This bidirectional relationship is thread throughout the discussions in the chapter. For example, Sections 3.53.7 examine the part of the committee’s charge related to “assessing recent and ongoing research in chemistry that draws upon chemistry’s unique capabilities in the synthesis, measurement, and modeling of molecular systems to advance QIS.” Sections 3.4 and 3.8 focus primarily on “assessing recent and ongoing research in QIS and advances in quantum information processing and technology that have the potential to transform various aspects of chemistry research.” This chapter also identifies current emerging tools and future needs in the area of instrumentation. Later, the chapter discusses the state of user facilities at different scales in the United States and recommends opportunities for advancing infrastructure in the near future.

3.1 DEVELOP NEW APPROACHES FOR ADDRESSING AND CONTROLLING MULTIPLE ELECTRON AND NUCLEAR SPINS IN MOLECULAR SYSTEMS

Several main categories of experimental approaches exist for addressing and manipulating electron and nuclear spins in molecules, including NMR, EPR, neutron scattering, and optical control/cycling (for quantum-state initialization and readout). This section discusses the advantages and limitations of each approach and outlines emerging areas of investigation with potential for further development.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

3.1.1 NMR and Molecular QIS

NMR has been widely employed by chemists in the molecular magnetism field since the initial demonstration of magnetic bistability in the Mn12-acetate single-molecule magnet (SMM; Sessoli, Gatteschi, et al. 1993), followed by the prized discovery of resonant quantum tunneling of magnetization in the same molecule (Hernandez et al. 1997). Soon after, the technique was applied more broadly in bulk samples, enabling fundamental insights into subtle effects that influence the quantum spin dynamics associated with a variety of different molecular magnetic clusters (Julien et al. 1999; Kubo et al. 2002; Micotti et al. 2006). Of particular importance are the works of Morello and others on the giant spin S = 10 Mn12 and Fe8 molecules (Baek et al. 2005; Chakov et al. 2006; Furukawa et al. 2001; Morello and de Jongh 2007; Morello et al. 2004) that led to an initial theoretical framework describing electron-nuclear decoherence processes associated with mesoscopic electron spin moments in molecules (Morello 2008; Stamp and Tupitsyn 2004). A feature common to almost all of the experimental NMR studies was a requirement that measurements be conducted to sub-kelvin temperatures on a wide variety of nuclei (1H, 55Mn, 19F, 53Cr, 57Fe), many of which are quadrupolar. As such, none of these studies are possible using commercial NMR instruments; they require strong collaboration between chemists and physicists as well as unique facilities such as the U.S. National High Magnetic Field Laboratory (NHMFL; Chakov et al. 2006). The need for unique facilities developed by multidisciplinary teams is a common theme in much of the research in this field.

Conventional NMR spectroscopy continues to find utility at the interface between chemistry and QIS as research has shifted toward the study of simpler monometallic systems and their potential use as spin qubits (Gimeno et al. 2021). For example, as illustrated in Figure 3-1, recent studies on a [YbIII(trensal)] molecule demonstrate coherent manipulation of the 173Yb (I = 5/2) nucleus at microwave frequencies typically associated with EPR (Hussain et al. 2018). Strong electron–nuclear hyperfine coupling together with a measurable nuclear quadrupolar interaction again necessitated the use of a specialized home-built broadband NMR spectrometer, HyReSpect (Allodi et al. 2005). The novelty of this work lies in the fact that the 73Yb nuclear qudit, with dimension d = 6, can encode the effective S = ½ electronic moment of the YIII ion with embedded basic error protection. Further development of ultra-broadband instrumentation to enable the application of the most advanced frequency-swept NMR methods (Altenhof et al. 2019) to this area of study is important, as recent EPR investigations (Kundu et al. 2022) highlight cases with nuclear zero-field quadrupole splitting approaching magnitudes typically associated with electrons (i.e., in the gigahertz range). Meanwhile, elegant synthetic work involving substitutional patterning of 1H and 79/81Br

Image
FIGURE 3-1 (left) Multilevel structure of the [YbIII(trensal)] molecule consisting of (center) a coupled electronic doublet (effective S = ½) and nuclear sextet (I = 5/2). A sizable nuclear quadrupole interaction allows selective excitation of individual nuclear transitions. (right) The multilevel structure of this qudit could be exploited to encode and operate a qubit with embedded basic quantum error correction, with fast gate (Rabi) operation times due to strong electron-nuclear mixing.
SOURCE: Hussain et al. 2018.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

nuclear spins on polybrominated catechol ligands and various complexes formed by coordination of the ligands to diamagnetic TiIV (Johnson, Jackson, and Zadrozny 2020) and magnetic VIV (S = ½) (Jackson et al. 2019) ions reveals a pronounced modulation of the proton nuclear spin dynamics in the ligand shell. The VIV example shows how the 1H dynamics and, hence, the nuclear spin patterning can ultimately influence the electronic coherence, suggesting important design principles for molecular spin qubits (see Figure 3-2).

Beyond conventional NMR spectroscopy, several related resonance techniques show tremendous promise for interrogating nuclear spins in areas of interest to chemists working in the QIS field. Several electron and electron-nuclear double resonance (ENDOR) spectroscopies are described in later sections of this report (Goldfarb 2017; Greer et al. 2018; Harmer 2016; Kundu et al. 2023; Sato et al. 2007, 2009; Van Doorslaer 2017; Wang et al. 2018). A particular advantage of these methods is that they often inherit the far greater sensitivity of EPR, in comparison to NMR, and, in some cases, are capable of achieving extreme broadband sensitivity to a multitude of nuclei in a single experiment. Meanwhile, the sensitivity of NMR can also be enhanced greatly by means of dynamic nuclear polarization (DNP; Barnes et al. 2008), whereby electron polarization is dynamically transferred to nuclei via EPR excitation. Although most current applications of DNP-enhanced NMR are in the biophysical research arena (Can, Ni, and Griffin 2015), one can envision a wide range of areas where this blossoming technique can impact research in molecular QIS. First and foremost, DNP enhancement mechanisms (of which there are many) depend intimately on the nature of the coupling between electron and nuclear spins. Thus, DNP-NMR studies can provide valuable insights into decoherence processes associated with molecular spin qubits. Meanwhile, studies have shown that one can use DNP to initialize or control the nuclear environment, leading to a range of intriguing possibilities including massive electron-nuclear entanglement (Simmons et al. 2011), enhanced electronic coherence (Bluhm et al. 2010), and even execution of multiple electron–qubit gate operations (Foletti et al. 2009). While these examples have so far been limited to solid-state semiconductor platforms, application to the molecular QIS field is ripe for investigation. However, this will require cross-fertilization of ideas between chemists working on DNP-NMR and synthesis of molecular qubits, as well as wider access to state-of-the-art magnetic resonance instrumentation and expertise. Finally, DNP offers tremendous prospects for applying NMR to the study of molecules on surfaces (Rossini et al. 2013), where it can provide a means to overcome sensitivity limitations due to low areal concentrations. As illustrated in later sections of this chapter, the study of molecules deposited on solid substrates is an important and growing area of investigation, with potential for the development of scalable device architectures.

Image
FIGURE 3-2 Nuclear spin-pattern control of the phase memory time in a series of V(IV) complexes. (left) Protons at symmetry equivalent sites have identical chemical and paramagnetic shifts, thus promoting resonant nuclear spin dynamics that cause decoherence of the V(IV) electron spin. (right) Temperature dependence of the electronic phase memory time, Tm, for two complexes with protons patterned at symmetric and asymmetric positions, demonstrating the significantly reduced coherence in the former.
SOURCE: Jackson et al. 2019.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

3.1.2 CW EPR and Molecular QIS

EPR has been applied extensively by chemists (and physicists) in the area of molecular quantum spin science, dating back to the very first studies of the Mn12 SMM (Gatteschi et al. 2006; Sessoli, Tsai, et al. 1993). Of particular note is the crucial role EPR played in achieving an understanding of the QTM phenomenon (Barra, Gatteschi, and Sessoli 1997; Barra et al. 2007; del Barco et al. 2005; Hill 2013; Hill, Edwards, Jones, et al. 2003; Wilson et al. 2006). As is the case with NMR, the instruments employed in these investigations are often found only in the laboratories of a few expert investigators or at national facilities (Baker et al. 2015). Moreover, even though targeted QIS applications are likely to employ low-field magnets and frequencies matching current communications bands, fundamental spectroscopic investigations often require both high magnetic fields that sometimes exceed those attainable in commercial magnets (Feng et al. 2012; Marriott et al. 2015; Ruamps et al. 2013; Zadrozny et al. 2012) and specialized high-frequency millimeter-wave to terahertz sources and detectors. In particular, the strong spin–orbit coupling and large spin values (giant spin in some polynuclear cases; see Nehrkorn et al. 2021) associated with transition metal and lanthanide complexes can give rise to extremely broad spectral splitting patterns. In comparison to NMR, where chemical shifts are usually measured in parts-per-million, so-called zero-field splitting (ZFS) in EPR can often significantly exceed the unperturbed Larmor frequency (28 GHz/T). This has spurred the development of ultra-wideband EPR instruments that span the millimeter, terahertz, and infrared (IR) spectral ranges (Hassan et al. 2000; Mola et al. 2000; Takahashi and Hill 2005; van Tol, Brunel, and Wylde 2005), including unconventional frequency domain approaches (Blackaby et al. 2022; Hay et al. 2019; Neugebauer et al. 2018; van Slageren et al. 2003) employing synchrotron sources in some cases (Schnegg et al. 2009; Suturina et al. 2017). By comparison, most commercial spectrometers are limited to frequencies below 10 GHz (~3 cm), with a few high-end instruments operating at 34, 94, and 263 GHz.

While much of the QIS-related EPR is currently performed using transient spectrometers (see Section 3.1.3), CW measurements have provided important fundamental chemical insights into this area of investigation in several cases. Motivated by the early quantum computing proposal by Loss and DiVincenzo (1998) involving exchange-coupled quantum dots, several groups investigated engineering coherent interactions within pairs (dimers) of magnetic molecules. The first example involved [Mn4]2 dimers, where weak hydrogen bonding interactions were crystallographically imposed (Hill, Edwards, Aliaga-Alcalde, and Chrisou 2003). Subsequent synthetic refinements gave rise to covalently linked [Mn3]2 supramolecular dimers that were shown to retain their coherent quantum properties in dilute frozen solutions (Ghosh et al. 2021; Nguyen et al. 2015), thereby demonstrating that the dimers retain their properties outside of a crystal. Meanwhile, work on {Cr7Ni} qubits demonstrated redox switchable interactions through triangular {Ru2M} linker molecules (M = Zn, Ni, or Co) (Figure 3-3; Ferrando-Soria, Magee, et al. 2016), suggesting possibilities for local switching via a gate electrode or a scanning tunneling microscope (STM) tip. More recent studies have focused on asymmetric lanthanide dimers (Giansiracusa et al. 2018; Luis et al. 2020).

3.1.3 Transient EPR

One of the earliest descriptions of a potentially realizable quantum computer considered localized electronic states in polymers (Lloyd 1993), followed soon after by a proposal to implement Grover’s search algorithm using the Hilbert space associated with a giant-spin polynuclear transition metal cluster (Leuenberger and Loss 2001). Common to both of these proposals is the essential role of pulsed EPR, which is the method of choice for coherent quantum manipulation of electron spins in molecules (Sato et al. 2009; Wolfowicz and Morton 2016).

Coherent pulsed EPR experiments have been used to obtain structural details of biologically relevant transition metal complexes as far back as the 1980s (Thomann et al. 1987). However, Ardavan and colleagues (2007) first asked whether coherence times in molecular magnets would be long enough to permit quantum information processing, reporting a phase memory time of several microseconds for a deuterated spin S = ½ {Cr7Ni} wheel complex. This marked the dawn of an era that extends to the present day; synthetic inorganic chemists have systematically explored the factors influencing relaxation times in molecular magnets (Gaita-Ariño et al. 2019), with ever-increasing T2 and Tm being reported from liquid helium temperatures all the way to room temperature

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-3 (top) Dimers of spin S = ½ {Cr7Ni} qubits linked by redox switchable {Ru2M} oxo-centered triangles (M = 3d metal, not shown), enabling chemical control of the coupling between the qubits. This control is demonstrated by cyclic voltammetry (lower left) and electron paramagnetic resonance (EPR) (lower right); in the latter case, one observes the coupling via a splitting of the EPR spectrum that can be turned “on” and “off” chemically.
SOURCE: Ferrando-Soria, Magee, et al. 2016.

(Figure 3-4; Atzori et al. 2016; Bader et al. 2014; Bertaina et al. 2008; Warner et al. 2013; Zadrozny et al. 2015). The vast majority of these investigations have been carried out using high-end commercial EPR spectrometers operating in the X-band frequency range (9–10 GHz) down to liquid helium temperatures. One or two studies have been carried out at lower frequencies using both commercial (Zadrozny et al. 2017) and home-built (Collett et al. 2019) instruments. These concerted activities have significantly advanced the understanding of spin-lattice (Kazmierczak, Mirzoyan, and Hadt 2021; Kragskow et al. 2022; Lunghi and Sanvito 2020) and spin-spin (Canarie, Jahn, and Stoll 2020; Chen et al. 2020; Kundu et al. 2023) relaxation in magnetic molecules, leading to the various strategies for enhancing coherence that were described in detail in Chapter 2 of this report: minimization of

Image
FIGURE 3-4 Some landmarks in the development of molecular complexes for quantum technologies. Timeline for some of the most relevant spin qubits made from transition metal ions in terms of quantum coherence time (T2).
SOURCE: Gaita-Ariño et al. 2019.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-5 (left) Molecular structure of the nuclear spin-free [Cr(C3S5)3]3− complex: pink, yellow, and gray spheres represent Cr, S, and C atoms, respectively. (center) Calculated splitting of the MS energy levels in a 2000 G magnetic field applied along the z axis of the molecule. The circular arrows illustrate formally allowed (red) and forbidden (blue) transitions within the spin S = 3/2 CrIII manifold. (right) Rabi oscillations corresponding to one of the formally forbidden transitions.
SOURCE: Fataftah et al. 2016.

spin–orbit coupling (Ariciu et al. 2019; Graham et al. 2014), suppression of spin–vibrational coupling through the design of ligand field symmetry (Kazmierczak, Mirzoyan, and Hadt 2021), use of nuclear spin-free metals (Figure 3-5; Bader et al. 2017; Fataftah et al. 2016) and ligands (Yu et al. 2016), use of nuclear spin patterning (Jackson et al. 2019), and use of clock transitions (Collett et al. 2019; Gaita-Ariño et al. 2019; Harding et al. 2017; Kundu et al. 2022, 2023; McInnes 2022; Shiddiq et al. 2016; Zadrozny et al. 2017).

As with CW EPR, metal-based spin qubits frequently necessitate measurements at frequencies considerably higher than the X-band range. In most cases, this is due to the combined influences of spin–orbit coupling and the ligand field, which gives rise to a large ZFS of the 2S + 1 spin eigenstates, for S > ½ (Fataftah et al. 2020; Takahashi et al. 2009). Even for the S = ½ case, a recent lanthanide example demonstrates that the electron-nuclear hyperfine interaction is of a sufficient magnitude that the spectrum cannot be resolved in a commercial X-band instrument (Kundu et al. 2022). Measurements at higher frequencies are challenging due to the lack of available microwave sources with sufficient power to achieve short spin rotation pulses. Although very high-end commercial spectrometers operating at higher frequencies and magnetic fields are available, they have found little utility in this area of research, primarily due to a lack of power and a lack of flexibility for studying metals with broad EPR lines. One notable exception is the case of an Fe4 molecule with a zero-field excitation that exactly matches the 94 GHz operating frequency (W-band) of a commercial instrument, thus enabling coherent spin manipulation in zero field (Schlegel et al. 2008). However, most if not all other notable high-field/frequency transient EPR studies have been conducted using custom-built hardware that is described in later sections of this chapter. An important early finding from these studies was a significant enhancement in coherence brought about by polarization of the electron spin bath at pumped liquid helium temperatures (Takahashi et al. 2008, 2009, 2011; Wang et al. 2011) (see Figure 3-6). A frequency of 240 GHz is equivalent to a thermal energy of ~11.5 K. Hence observing an EPR signal at low temperatures necessarily involves excitation from the ground spin projection state that is isolated from the first excited state by 11.5 K. Therefore, one can achieve an electron spin polarization exceeding 99.9 percent at a temperature of 1.5 K (pumped 4He). As a result, electron spin-spin fluctuations are almost completely suppressed, and one can observe coherent electron spin dynamics in a highly concentrated crystalline sample, as has been demonstrated for several polynuclear Fe clusters (Takahashi et al. 2011; Wang et al. 2011).

Besides measuring T1 and demonstrating the possibility for coherent manipulation of spin states (e.g., Rabi oscillations) in a range of exotic molecular spin systems (see, e.g., Gould et al. 2021; Hu et al. 2018), pulsed

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-6 Temperature dependence of the spin-spin relaxation time, T2, for two samples (red and green data) of the Fe8 molecular magnet, with the field applied at two orientations [(a) //x and (b) //y] in the hard plane. The thick blue lines are calculations including phonon and magnon contributions. Insets: partial contributions to 1/T2 calculated from magnons (dashed line) and phonons (long-dashed line). The scale on the right side of the main panel indicates the decoherence Q-factor.
SOURCE: Takahashi et al. 2011.

EPR provides access to a wealth of microscopic information via coherent multidimensional pump-probe techniques. In its most basic form, pulsed electron-electron double resonance (ELDOR) involves excitation at one frequency followed by a coherent detection (often by a standard π/2 – τ – π – τ – Hahn echo sequence) at a second frequency within an inhomogeneously broadened EPR spectrum. Such experiments can, for example, provide information on electron spectral diffusion by monitoring changes in polarization at one location/frequency in the spectrum brought about by saturation at another (Wang et al. 2018), thus yielding microscopic insights into electronic relaxation.

A fully coherent pulsed ELDOR sequence, also known as double electron-electron resonance (DEER), involves the execution of an inversion pulse at a frequency ν2, during either a two-pulse Hahn echo sequence or a three-pulse sequence with a second refocusing π pulse at a frequency ν1 (Jeschke 2018). This technique, which was developed for structural studies of biomolecules, precisely probes the weak dipolar coupling between two spin labels by monitoring changes in the coherence of one of the spins (at frequency ν1) by inverting the polarization of the other spin (at frequency ν2) as a function of delay time, t, within the echo sequence. The result is a modulation of the echo intensity, the Fourier transform of which corresponds to the dipolar coupling in frequency units. The coherent nature of the DEER technique enables the measurement of extremely weak dipolar couplings (~MHz) that would otherwise be buried (i.e., unresolvable) deep within the inhomogeneously broadened CW EPR spectra. Increased spectral resolution at high magnetic fields further permits enhanced orientation selectivity through excitation and detection at spectral locations corresponding to different components of the electron g-tensor. In this way, one obtains information not only on the distances between spins but also on the relative orientations of their g-tensors (Stevens et al. 2016). As depicted in Figure 3-7, DEER has recently been deployed to measure electron–electron coupling within dimers of {Cr7Ni} spin qubits that have been proposed as molecular two-qubit gates (Ardavan et al. 2015), representing a first step toward the demonstration of electron quantum logic in molecules. Indeed, several proposals exist in the literature centered on the use of asymmetric dimers (or oligomers) as two-qubit gates (Aguilà et al. 2014; Collett et al. 2020; Ferrando-Soria, Moreno Pineda, et al. 2016; Uber et al. 2017; Ullah et al. 2022). The asymmetry is necessitated by the requirement that the spins be spectrally distinguishable. However, the demonstration of universal two-qubit logic gates remains a major outstanding goal, for which the primary obstacle is the limited bandwidths (~1 GHz) of current state-of-the-art transient EPR spectrometers (i.e., the limited ability to excite at two well-separated frequencies; Cruickshank et al. 2009). This issue is discussed further in Section 3.11.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-7 (left) 2.5 K double electron-electron resonance (DEER) time traces for a dimer of {Cr7Ni} rings connected by rigid organic thread molecules that contain amine binding sites (see Figure 3-4). (center) Different ring orientations relative to the external magnetic field can be excited by the frequencies of the DEER pulses owing to g-factor anisotropy (red: high excitation, blue: low excitation). (right) Fourier transforms of the time-domain data: open circles show raw data; solid lines indicate filtered data. The principal difference between the unfiltered data and the filtered data is the absence in the latter of a component at ~18 MHz arising from the coherent coupling to protons (electron-spin-echo envelope modulation).
SOURCE: Ardavan et al. 2015.

Pulsed EPR can also provide access to electron-nuclear hyperfine and superhyperfine (i.e., coupling of electron density on one atom to nuclear spin on coordinating atoms) interactions. The simplest one-dimensional example is the electron-spin-echo envelope modulation (ESEEM) effect, in which the excitation of formally forbidden electron-nuclear transitions (e.g., the zero-quantum [ZQ], uD → dU; and double-quantum [DQ], dD → uU, resonances, where U and D correspond to up and down and lower- and uppercase denote the nuclear and electron spins) during a two- or three-pulse sequence results in temporal modulations of the echo intensity as a function of delay time on account of the change in nuclear polarization during the coherent sequence (Van Doorslaer 2017). The Fourier transform of the modulation provides information on the coupling of the electron spin to surrounding nuclei; this method has been used recently to demonstrate electron–nuclear decoupling at a clock transition (Kundu et al. 2023). A 2D version of ESEEM—so-called hyperfine sublevel correlation spectroscopy—provides an unprecedented resolution of hyperfine and quadrupolar interactions. This technique is also useful in the study of molecular spin qubits (Ariciu et al. 2019; Atzori et al. 2018), enabling assessment of the degree of electron delocalization from metal to the coordinating ligands.

Multifrequency pulsed ELDOR measurements also provide direct information on NMR frequencies (Goldfarb 2017)—so-called ELDOR-detected NMR—again through excitation of formally forbidden ZQ and DQ electron-nuclear transitions. These transitions become weakly allowed when the electron-nuclear hyperfine interaction breaks rotational symmetry, as is the case, for example, for dipolar coupling. In a 2D experiment, a saturating pump pulse at frequency ν2 burns multiple holes in the EPR spectrum, with the ZQ and DQ transitions shifted with respect to the detection frequency, ν1, by an amount Δν = ±γnB0 ± ½A, where γn is the gyromagnetic ratio of the coupled nucleus, B0 the applied magnetic field strength, and A the hyperfine coupling strength. The major advantage of this technique is that it can provide very wideband (in comparison to conventional NMR) information on the electronic coupling to multiple distinct nuclei in a single experiment, while also inheriting the superior sensitivity of EPR. Again, the method is useful in the study of molecular magnets (Figure 3-8), providing information on electron delocalization in a metal-metal bonded dinuclear Fe-V complex (Greer et al. 2018). In a

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-8 A two-dimensional color plot of the experimental electron-electron double resonance (ELDOR)-nuclear magnetic resonance (NMR) spectra of an Fe-V metal-metal bonded molecular complex; red and blue correspond to high and low intensities, respectively. The right panel shows the echo-detected electron paramagnetic resonance spectrum recorded at 8 K and 94 GHz. The top panel contains a single ELDOR-NMR trace recorded at ~3.19 T. The isotropic transition centered at Δν ~ 20 MHz accounts for the interaction between the unpaired electron spin density and the deuterated solvent. The axial features centered at Δν ~ 35 MHz arise from the V hyperfine interaction; in the weak-interaction limit, they are centered at the 51V nuclear Larmor frequency and then split by the magnitude of the hyperfine coupling, which is anisotropic (different splitting for excitation at different components of the g-tensor).
SOURCE: Greer et al. 2018.

study conducted at the W-band on a custom-built high-power spectrometer, Cruickshank and colleagues (2009) demonstrate the advantages of increased orientation selectivity at high fields, enabling correlation between the electron g- and hyperfine coupling tensors. Substitution of one of the ELDOR pulses with a radiofrequency (RF) NMR pulse leads to the ENDOR technique (Harmer 2016), which is useful at the interface between chemistry and QIS (Lutz et al. 2013; Sato et al. 2009; Weiden, Käss, and Dinse 1999). When combined, these multidimensional pump-probe EPR techniques provide access to complementary information about electron-electron and electron-nuclear spin-spin interactions, from which one may infer exquisite microscopic details of physical and electronic structures and spin relaxation behavior. However, more widespread deployment of these methods has likely been limited by the lack of available state-of-the-art transient EPR spectrometers and by the unfamiliarity with advanced (i.e., complex) EPR methodology among the synthetic chemistry community.

3.1.4 Neutron Scattering

The key advance in neutron scattering over the past decade has been the development of a new generation of high-flux cold-neutron time-of-flight spectrometers. They are equipped with arrays of position-sensitive detectors that enable efficient measurement of neutron scattering cross sections as a function of energy and the three components of the momentum transfer vector Q, and in vast portions of reciprocal space (Chiesa et al. 2017). The availability of large single crystals of magnetic molecules then permits very detailed four-dimensional (4D) inelastic neutron scattering (INS) spectroscopy, providing unprecedented insight into coherent spin dynamics (Figure 3-9). Indeed, 4D INS enables complete evaluation of the dynamical correlation function, S(Q,ω), without reliance on any spin Hamiltonian model.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-9 (top) Illustration of the experimental setup, with incident neutrons scattered from a Mn12 molecule into an array of position-sensitive detectors that provide four-dimensional (4D) energy and momentum resolution. (bottom) Snapshot of the 4D inelastic neutron spectrum of Mn12, as a function of the transferred energy E and momentum Q. Here, the scattered neutron intensity is shown as a function of Qx, Qy, and E, integrated over the full measured Qz range.
SOURCE: Garlatti et al. 2019.

These new INS capabilities are starting to impact research in the molecular QIS sphere in dramatic ways. Recent examples include direct evidence for interactions between electronic (i.e., spin) and vibrational degrees of freedom associated with crystals containing vanadyl spin qubits, thus providing crucial information related to spin–phonon coupling strengths (Garlatti et al. 2020), and portrayed entanglement within dimers of {Cr7Ni} molecular spin qubits (Garlatti et al. 2017). One may expect further applications of this powerful new capability over the next 10 years, albeit such measurements are limited to just a handful of large-scale neutron scattering facilities around the world.

3.2 ENHANCE OPTICAL CONTROL IN MOLECULAR SYSTEMS IN QIS

The diversity of molecular species, coupled with the precision available with from molecular synthesis techniques, offer the prospect of building molecular qubits that are uniquely tailored for applications ranging from quantum networks to quantum-enhanced sensing (Wasielewski et al. 2020). As discussed in Chapter 2, the strong interaction of molecules with light (visible to near-infrared [NIR] radiation) in optically active molecules offers a powerful means of control. Molecules with optically addressable spins allow high-fidelity initialization and detection of qubit states (Awschalom et al. 2018). Coherent coupling of molecules with photons allows for long-range entanglement of qubits, opening the possibility of quantum networks over long distances. Separately, photons can be used to transduce quantum information to other physical platforms. Visible/NIR also offers spatial information down to the length scale of the optical wavelength, and potentially below that with the use of near-field techniques.

3.2.1 Optical Quantum-State Initialization

The large number of degrees of freedom (e.g., vibrational, rotational, electronic spin, nuclear spin) present in a molecule or molecular system means that at room temperature, many quantum states are energetically

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

accessible and populated via the Boltzmann distribution. The task of initializing a molecule into a single quantum state involves removing entropy associated with the Boltzmann distribution from a molecular system. Coupling molecular systems to light offers a convenient way to accomplish this: entropy can be transferred to light that propagates away from the system. If this process occurs on a timescale much shorter than that of thermalization with the environment, high-fidelity initialization of molecular qubits can occur even at room temperature, which is technologically desirable.

For optical initialization to occur, one first requires optically active molecules—those with high oscillator strength for an electric dipole transition between the ground and some optically excited state at technologically accessible wavelengths. Second, a configuration where certain quantum states are “dark” to incoming light is required. The molecule–light system must be engineered into such a configuration. When this occurs, repeated scattering of incoming photons “pump” molecules into the desired dark quantum state, a technique known as optical pumping. Concretely, the “darkness” required for internal state initialization can be achieved through spectral separation of optical transitions, or through control of the light polarization. Examples can be found in solid-state qubits formed by nitrogen-vacancy (NV) centers in diamond and chemically designed optical Cr+ spin qubits. The relevant qubit (“spin”) states in the ground electronic manifold are split by a large energy difference, allowing qubit states to be spectrally resolved (Awschalom et al. 2018; Jelezko and Wrachtrup 2006; Rodgers et al. 2021). By addressing a specific spin state with light, a qubit can be optically pumped into the unaddressed “dark” states. This form of optical pumping is also known as spectral hole burning. In gas-phase molecules that permit optical cycling (e.g., optical cycling centers [OCCs]), arrays of single diatomic molecules have been initialized through a polarization-sensitive optical pumping scheme (Holland, Lu, and Cheuk 2022).

3.2.2 Optical Quantum-State Readout

By repeatedly exciting a molecular qubit optically and collecting the resulting laser-induced fluorescence, the qubit can be detected. Quantum-state resolution is attained when the optical transitions to different spin states differ in energy and color. Through state-selective optical excitation, isolated OCCs individually trapped in optical tweezer arrays have been detected with quantum-state resolution and with high fidelity (Figure 3-10; Anderegg et al. 2019). In addition to the frequency of the emitted fluorescence, in certain systems, the polarization of an emitted photon can be entangled with the qubit state. This can be harnessed to create quantum networks of distant molecular qubits that are entangled, which can be a resource for quantum communication (Togan et al. 2010).

Image
FIGURE 3-10 (a) Schematic of optical tweezer array setup used to trap individual diatomic optical cycling centers (OCCs), and an average fluorescent image of single OCCs in a five-qubit array. (b) Single-shot fluorescent images of defect-free arrays of diatomic OCCs (CaF) held in a reprogrammable array of optical tweezer traps.
SOURCES: (a) Anderegg et al. 2019; (b) Holland, Lu, and Cheuk 2022.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

3.2.3 Optical Cycling in Molecular Systems for QIS Applications

Both for optical pumping and quantum-state readout via molecular fluorescence, it is desirable to be able to repeatedly scatter photons off a molecule (i.e., optically cycle photons). For quantum-state initialization through optical pumping, optical cycling is needed since the pumping process can be stochastic, and multiple excitations and decays are needed for a molecule to fall into the desired dark state. For fluorescent quantum-state readout, typical collection efficiencies are much less than unity, implying that multiple photons need to be emitted from a molecule for high-fidelity detection. Achieving optical cycling depends on finding an optically closed system; this can be accomplished by using optically active (high oscillator strength) molecular qubits with atom-like structures or by finding molecules with particularly favorable cycling properties.

Optical cycling, optical initialization of molecular bits, and optical quantum-state (spin) readout have been accomplished in a variety of molecular systems including molecular spin qubits and a variety of gas-phase molecules. A common challenge is decoupling the optical excitations and decays from exciting molecular vibrations.

When a molecule is optically excited, the subsequent decay can be accompanied by a vibrational excitation. That is, the molecule can fall into a different vibrational state that is off-resonant from the initial optical excitation. This problem becomes more severe as the size of a molecule grows and the number of vibrational modes increases. As described in the subsequent section, recent work has shown that this problem can be circumvented in OCCs, in which a very high degree of optical cycling can be achieved.

To provide a point of comparison, the problem of vibrational branching in molecules has an analog in optically active solid-state defects such as NV centers. In these systems, the emission of a photon from an optically excited defect can be accompanied by lattice excitations (phonons), resulting in an emission spectrum containing a narrow line (zero-phonon line [ZPL]) accompanied by a broad background corresponding to various phonon excitations. The latter can be problematic in quantum applications that require indistinguishable photons such as quantum teleportation and quantum networks of entangled spins (Hensen et al. 2015). Owing to conservation of energy, photons emitted in the ZPL can be distinguishable in color from phonon-assisted emission. Even for measurements short enough that the frequency difference is not a concern, phonon-assisted emission can destroy quantum entanglement because the accompanying phonons can lead to rapid decoherence.

3.2.4 Optical Cycling Centers

An area emerging at the intersection between atomic molecular and optical (AMO) physics and chemistry is the identification and control of gas-phase molecules with favorable optical cycling properties, known as OCCs. In the field of AMO physics, full quantum control over molecules at ultracold (< mK) temperatures has been pursued intensely in the past two decades, primarily because of the promise of molecules as a new platform for quantum science. Motivated by the high degree of control achievable in atoms via laser light, initial approaches focused on coherently assembling a molecule out of two ultracold atoms (Ni et al. 2008, 2009; Ospelkaus et al. 2010), whereby one inherits full quantum-state control available with atomic techniques. Although this has been a fruitful path and has led to the observation of reactions modulated by quantum exchange statistics (Ospelkaus et al. 2010) and quantum state–resolved chemistry (J. Liu, Mrozek, et al. 2021; Y. Liu et al. 2021; Liu and Ni 2022), the method of coherent molecular assembly from atoms is difficult to extend beyond diatomic molecules.

An alternative approach generalizable to a large class of molecules has also been pursued. Rather than assembling molecules from their constituent atoms, physicists have sought to directly adapt optical control techniques developed for atoms, such as optical pumping and laser cooling, to molecules. These require optical cycling at a high level of photons (104 to 105), a task that appeared difficult initially. But with key developments in identifying optically cyclable molecules and devising an optical cycling scheme based on rotational selection rules (Di Rosa 2004; Stuhl et al. 2008), optical cycling of 104 to 105 photons enabling laser cooling has now been achieved for many gas-phase diatomic molecules (Anderegg et al. 2017; Barry et al. 2014; Collopy et al. 2018; Shuman et al. 2009; Truppe et al. 2017). In these diatomic molecules, the ability to optically cycle has led to many advances that pave the way toward full quantum control. The advances enabled by optical cycling include new techniques to cool and trap large samples of molecules (Anderegg et al. 2018; McCarron 2018),

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

coherent quantum control of rotational states (Williams et al. 2018), and nondestructive fluorescent detection of ultracold molecules (Cheuk et al. 2018).

Notably, with relevance to quantum information processing and quantum simulation, recent work with laser-cooled diatomic molecules has demonstrated high-fidelity detection of arrays of single molecules held in optical tweezer traps (Anderegg et al. 2019; Holland, Lu, and Cheuk 2022). Recent work with molecular tweezer arrays has even demonstrated high-fidelity positioning and quantum-state initialization of single molecules within a large array (Holland, Lu, and Cheuk 2022). Very recently, on-demand entanglement between two molecules spatially separated on the micron scale was demonstrated for the first time, long coherence times on the 100 ms timescale were achieved, and a two-qubit gate sufficient for universal quantum computation was implemented in these arrays (Bao et al. 2022; Holland, Lu, and Cheuk 2022). These demonstrations establish the building blocks needed for quantum information processing and quantum simulation, and make molecular tweezer arrays a promising new platform for quantum science. Looking ahead, the platform of tweezer arrays with optically cyclable molecules could allow high-fidelity individual qubit addressing. By leveraging recent advances in optical tweezer arrays filled with single atoms, these molecular arrays could also be scalable in the near term to hundreds of qubits, as has been demonstrated in their atomic analogs (Ebadi et al. 2021; Scholl et al. 2021).

In the past few years, new work extending beyond diatomic radicals has explored whether optical cycling can be applied to more complex molecules and where new chemical principles can be found that guide the design of these molecules. On the physics front, the search for large molecules with favorable optical cycling properties is motivated by the fact that certain molecules are particularly sensitive to the potential for new fundamental physics beyond the standard model. If these same molecules can be optically controlled at the quantum level, they could offer significant improvements in fundamental physics searches, such as those that look for an electron electric dipole moment (Augenbraun et al. 2020) and ultralight dark matter (Kozyryev, Lasner, and Doyle 2021). The extension of similar techniques to larger molecules could open up possibilities including quantum-enhanced precision measurements and chemistry at ultracold temperatures with entangled matter.

The quest to identify molecules with favorable optical cycling properties along with the discovery of guiding chemical principles is where AMO physics and chemistry intersect. This is a rapidly evolving area. So far, a theme that has guided work on identifying OCCs successfully is the M-O-R motif, where an alkaline earth metal atom (M) is bonded to a halide (Kozyryev, Baum, Matsuda, Boerge, and Doyle 2016; Kozyryev et al. 2019). These systems behave as gas-phase M+ cation radicals. Optical excitations primarily involve orbitals localized on the metal atom that are minimally coupled to vibrational modes, thereby permitting optical cycling. Based on this principle, theoretical calculations and, in some cases, experiments have shown that more complex M—O—R (where O is oxygen, and R is the organic group) molecules can minimize the issue of vibrational branching and permit optical cycling (Dickerson et al. 2022; Dickerson, Guo, Shin, et al. 2021; Dickerson, Guo, Zhu, et al. 2021; Ivanov et al. 2020; Kłos and Kotochigova 2020; Kozyryev, Baum, Matsuda, and Doyle 2016; Mitra et al. 2022; Zhu et al. 2022). Experimental work has demonstrated optical cycling in the M-O-R polyatomics CaOH and CaOCH3 (Hallas et al. 2023; Mitra et al. 2020; Vilas et al. 2022). Notably, the optical control over CaOH is rapidly approaching that of diatomic OCCs.

For even larger molecules, recent work has discovered new chemical principles that determine optical cycling properties. Theoretical work has shown that Franck–Condon factors in an M-O-R OCC can be tuned through chemical substitution in the R-ligand organic functional group (Figure 3-11; Dickerson, Guo, Shin, et al. 2021). Recent experiments have also verified that OCCs attached to aromatic groups can retain good cycling properties (Mitra et al. 2022; Zhu et al. 2022). Together, these works on M-O-R OCCs have also revealed a possible new chemical principle regarding optical cycling properties—the electron-withdrawing potential of the R ligand correlates well with the molecule’s electronic excitation energies and vibrational branching ratios.

The rapidly developing area of large OCCs has also increasingly emphasized the perspective of large OCCs as molecules functionalized with an optical cyclable quantum functional group. Viewing an OCC as a quantum functional group serving as a generic qubit moiety, there are proposals to attach optically controllable qubits to larger molecules and to surfaces (Guo et al. 2021; Zhu et al. 2022). If successful, these endeavors could open up new possibilities such as transducing information among several attached OCCs joined to a large molecule or surface and using attached OCCs as quantum sensors to probe the dynamics of their host molecules (Zhu et al. 2022).

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-11 (a) Diagram illustrating optical cycling centers (OCCs) built upon the M-O-R motif. A metal atom (Ca) is ionically bounded to a molecular fragment (phenyl ring) that can be tuned via chemical substitution (at the R positions). (b) Computed Franck–Condon factors of M-O-R OCCs of Ca-phenoxide and Sr-phenoxide with substitutions. The Franck–Condon factors determine the vibrational branching ratio when the molecule is excited from the ground state (X) to an excited state (A or B). (c) Computed Franck–Condon factors are shown as a function of Hammett’s total of the R substituents for CaOR (blue) and SrOR (red) species, illustrating the principle of modulating optical properties of OCCs via chemical substitution. The Hammett total indicates the electron withdrawing strength of the substituent functional group.
SOURCE: Dickerson, Guo, Shin, et al. 2021.

3.3 DEVELOP TECHNIQUES TO PROBE MOLECULAR QUBITS AT COMPLEX INTERFACES TO INFORM THEIR SYSTEMATIC CONTROL

3.3.1 State of the Art in Single-Molecule and Surface NMR

The need to obtain microscopic information concerning nuclear spin dynamics and structural details associated with molecules deposited onto surfaces has driven the development of several novel nuclear resonance methods. For example, 57Fe Mössbauer spectroscopy (sometimes dubbed nuclear gamma resonance), which has long been used in the investigation of iron-based SMMs (see, e.g., Zadrozny et al. 2013), has recently been applied to a monolayer film of an Fe4 molecular cluster (Cini et al. 2018). Measurements were performed on a 95 percent 57Fe enriched sample using a synchrotron gamma radiation source at the European Synchrotron Radiation Facility. Remarkably, these studies could capture details of structural deformations that escaped detection by conventional synchrotron techniques such as X-ray magnetic circular dichroism. Moreover, the measurements provided information concerning the fluctuation timescale associated with the magnetic moments. Meanwhile, β-detected NMR involves implantation of a magnetically polarized radioactive isotope that undergoes β-decay. By subjecting the sample to a static magnetic field, one can effect a loss of this polarization upon application of RF pulses that are resonant with the Larmor precession of the implanted ions. Analogous to muon spin relaxation (μSR), which has also been applied to the study of molecular spin qubits (Baker et al. 2016), the polarization can be measured from the asymmetry in β-emission from the implanted ions. In a recent example involving Mn12 molecules grafted onto a Si substrate, 8Li+ ions were selectively implanted with different depth profiles based on their implantation energy (Salman et al. 2007). The resulting NMR lineshapes could be used to infer the local distribution of static magnetic fields at different depths relative to the surface, in turn informing on the magnetic properties of the Mn12 molecules. These experiments were performed at the Isotope Separator and Accelerator at TRIUMF in Canada.

One of the landmark results in the area of molecular QIS involves the implementation of Grover’s quantum search algorithm using a single I = 3/2 nuclear spin carried by a TbIII bis-phthalocyanine complex within a molecular transistor (Godfrin, Ferhat, et al. 2017). As illustrated in Figure 3-12, the method relies on the magnetic bistability of the J = 6 spin-orbital moment of the TbIII ion, which can only flip from “up” (↑) to “down” (↓) through quantum tunneling at the crossing point of the MJ = ±6 Zeeman levels (see Figure 3-12) (Godfrin, Thiele, et al. 2017; Thiele et al. 2014; Vincent et al. 2012). These are split into four sub-pairs due to hyperfine coupling to the I = 3/2 nucleus. Consequently, the electronic moment knows about the quantum state of the nucleus; when the electronic spin flips from “up” to “down,” it induces a corresponding jump in the differential conductance of the

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-12 (a) Artist’s view of a nuclear spin qubit transistor based on a single TbPc2 (where Pc is phthalocyanine) molecular magnet, consisting of a Tb3+ ion (pink) sandwiched between two Pc ligands (white) and coupled to source and drain electrodes. The four nuclear spin states of the Tb3+ (colored circles) can be manipulated with electric field pulses. (b) Zeeman diagram of the TbPc2 molecular magnet, showing the hyperfine split electronic spin ground-state doublet as a function of the external magnetic field. Avoided level crossings (colored rectangles) allow for tunneling of the electron spin. (c) Jumps in conductance read out the state of the Tb3+ nuclear spin during field sweeps. (d) Histograms of the conductance jumps. (e) Maximum visibility of the Rabi oscillations. (f) Resonance shape of the three nuclear qubit transitions.
SOURCE: (a and b) Thiele et al. 2014 (c); Godfrin, Ferhat, et al. 2017.

transistor. The location in the magnetic field of this jump, therefore, provides a convenient readout of the nuclear state of the molecule. One can then perform logic operations on the nuclear spin via the application of selective microwave NMR pulses (see Figure 3-13), with the final readout achieved via the transistor. A quality factor can be deduced from the ratio of T2/Tp (where Tp is the time taken to execute a p-pulse), which gives Q = (300 ms)/(0.12 ms) = 2,500, corresponding to a fidelity of about 99.9 percent. The next step would involve single-qubit gate randomized benchmarking which, to the best of our knowledge, has not been performed for a molecular system to date. These and other related studies (Jo et al. 2006; Perrin, Burzurí, and van der Zant 2015) clearly demonstrate the feasibility of realizing molecular-scale electronic devices for QIS.

Image
FIGURE 3-13 (a) Schematic of surface nuclear magnetic resonance (NMR) experimental setup. Ubiquitin proteins attached to the diamond surface are probed using a proximal quantum sensor consisting of a nitrogen-vacancy (NV) center electronic spin and its associated 15N nuclear spin. (b) 2H and (c) 13C NMR spectra (red points) recorded at magnetic fields of 2473 and 2457 G, respectively, with Gaussian fits superimposed (black curves). (d) Scalings of resonance frequencies with an applied magnetic field.
SOURCE: Lovchinsky et al. 2016.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Finally, one can turn the single-spin sensing problem on its head by deploying more mature detection platforms such as magnetic resonance force microscopy (MRFM) or the NV defect center in diamond acting as a quantum reporter (discussed in detail in Chapter 2), thereby pushing NMR sensitivity toward the single-spin limit. Nanoscale nuclear magnetic resonance imaging came first (Degen et al. 2009; Hemmer 2013; Mamin et al. 2013; Staudacher et al. 2013), followed shortly thereafter by single-proton detection (Müller et al. 2014; Sushkov et al. 2014). As shown in Figure 3-13, these advances now permit NMR studies of individual proteins (Lovchinsky et al. 2016) and nanoscale nuclear quadrupolar sensing (Lovchinsky et al. 2017), opening up entirely new areas of chemical imaging and spectroscopy (Aslam et al. 2017; Liu et al. 2022; Rugar et al. 2015).

3.3.2 Single-Molecule EPR

Signatures of individual molecular spins within an ensemble were first reported via optically detected EPR in 1993 (Köhler et al. 1993; Wrachtrup et al. 1993). However, the first direct EPR detection of a single localized electron spin was achieved at IBM using MRFM (Rugar et al. 2004), a precursor to the NMR studies described above. A major breakthrough that would eventually lead to the possibility of performing transient EPR on single surface spins (see below) involved the use of an STM both for manipulation and detection of individual Fe adatoms placed on a MgO film (Baumann et al. 2015). This method, which was again developed at IBM, built upon a considerable body of scanning tunneling spectroscopy work on individual surface atoms that mimicked many of the well-known properties of molecular nanomagnets (see, e.g., Heinrich et al. 2004; Loth et al. 2010). Consequently, this work received much attention among chemists working in the area of molecular QIS. The EPR-STM technique has continued to advance at a breathtaking pace, with a number of laboratories around the world developing similar capabilities. These methods hold tremendous promise for exploring next-generation qubit platforms, particularly at the interface between chemistry and QIS. For example, a recent CW EPR investigation published in Nature Chemistry explored the surface chemistry of FePc and FePc dimers (where Pc is phthalocyanine) adsorbed on a MgO surface (Zhang et al. 2022). In particular, these studies revealed that the exchange coupling within the FePc dimers depends strongly on molecular geometry, which is dictated by the available surface binding sites. Other groups have looked at organic radicals adsorbed on graphite (Durkan and Welland 2002) and alkali metal dimers on MgO (Kovarik et al. 2022).

As illustrated in Figure 3-14, more recent developments of the EPR-STM method have demonstrated that one can deploy the entire transient EPR toolkit (Chen, Bae, and Heinrich 2022; Willke et al. 2021; Yang et al. 2019). This enables coherent manipulation and readout of individual spins associated with atoms or molecules deposited onto an insulating (typically MgO) film supported on a metallic substrate. RF excitation is achieved either via alternating current (AC) modulation of the tunneling current generated by the STM (the precise mechanism of

Image
FIGURE 3-14 (a) Experimental setup for a pulsed electron spin resonance–scanning tunneling microscope (STM). The STM is equipped with a radiofrequency (RF) generator and an arbitrary waveform generator. A sequence of RF and direct current pulses is delivered to an oxygen-site Ti atom on MgO/Ag(100) via the STM tip. (b) Rabi oscillations of the Ti spin at different RF powers (VRF). (c) Hahn echo measurements of the Ti spin.
SOURCE: Chen, Bae, and Heinrich 2022.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

this driving mode remains somewhat unclear) or via a separate RF antenna close to the tip. Detection is achieved by employing a magnetized STM tip that generates a spin-polarized tunneling current, enabling a projective readout of the state of the spin located below the tip. Averaging then yields the composition of the wave function (i.e., ∣ψ〉 = a∣↑〉 + b∣↓〉). By employing different excitation schemes, one can carry out all of the well-known pulsed EPR schemes (e.g., Rabi nutation, Ramsey fringe measurements, Hahn-echo decay measurements of the Tm associated with a single spin, and so on). The technique, which has been pioneered in Andreas Heinrich’s group at Ewha Womans University in Seoul, Korea, obviously requires access to high-end STM instruments along with considerable technical expertise. Most demonstrations until now have involved individual surface adatoms as well as pairs of atoms. However, one recent study focused on the coherent control of the spin associated with a single FePc molecule (Willke et al. 2021), emphasizing the tremendous potential of this relatively new method for coherent control of individual molecular qubits. An added powerful advantage of this technique is the possibility of moving atoms and molecules around on surfaces (Loth et al. 2012), enabling a very deliberate design of multispin architectures for studies of the physics associated with relatively simple model spin Hamiltonians. However, a potential drawback is the lack of an obvious pathway to scalable architectures with the possibility to manipulate large numbers of qubits. Nevertheless, EPR-STM offers a remarkably powerful tool for studying the quantum dynamics of individual (or small numbers of) molecules on surfaces. Limited availability of such instruments has resulted in a relatively low number of publications until now. Groups in Europe (Korvarik et al. 2022; Yan et al. 2020) and Asia (Chen, Bae, and Heinrich 2022; Kawaguchi et al. 2023) are increasingly turning to this technique, even though it was first developed in the United States (Baumann et al. 2015). Obviously, less exotic (and less costly) EPR techniques should be used to pre-screen candidate molecules. However, limited access to EPR-STM platforms, particularly in the United States, is preventing research at the intersection of QIS and chemistry to move forward.

3.3.3 Chemical Tailoring of Solid-State Spin Defects

Defects in semiconductors have become powerful hosts for spin-based qubits in the solid state within materials including diamond, silicon carbide, and silicon. By exploiting defects generated via electron or ion radiation as well as naturally occurring defects in these materials, researchers have demonstrated precise quantum control of individual electron and nuclear spins, robust entanglement, electron-nuclear quantum registers, and increasing control of the spin-optical interface at the level of single photons. For example, like NMR, the use of single NV centers close to the surface of diamond nanocrystals or nanopillars to sense single molecular spins deposited on their surfaces shows tremendous promise for chemical EPR investigations. Recent work by Pinto and colleagues (2020) has demonstrated coherent readout and control of an endohedral 14N@C60 electronic spin via this method, including transient measurements. This approach can also be extended to nanoscale samples at high magnetic fields and frequencies (Fortman et al. 2020). In particular, NV-detected EPR offers the possibility to detect single electron spins and, therefore, to investigate biological processes at the single-molecule level. As such, NV-detected EPR with single-spin sensitivity can potentially eliminate ensemble averaging in heterogeneous systems (Giovannetti 2004).

Meanwhile, driven by predictive theoretical work, several defect-based states have been engineered for practical applications including quantum memories and optical emission in the telecom regime. Impressive quantum sensing of electric, magnetic, and strain fields has been shown along with dramatic improvements in small-volume and single-spin NMR. However, these material systems face considerable challenges for quantum science and engineering, including creating scalable and identical quantum states, mitigating the impact of strain and isotopic variations, as well as controlling decoherence-driven charge fluctuations from interfaces and unintentional dopants.

The selection of appropriate hosts and defects, and their mutual interactions with one another, remains both a challenge and an opportunity for QIS research. There have been considerable advances in understanding the mechanisms of spin decoherence in semiconductors and solid-state nanostructures. State-of-the-art synchrotron spectroscopies, coherent Bragg diffraction imaging, and high-resolution magnetic resonance techniques have proven to be powerful tools to improve the quality of host materials. In addition, improvements in first-principles approaches to understanding and predicting properties (e.g., density functional theory) have led to a deeper physical understanding of quantum decoherence and successful predictions in mitigating many sources of noise,

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

from dynamical decoupling to unique pulse sequences. In addition, there have been successful efforts aimed at harnessing the capabilities of today’s electronics technologies for quantum-state control, including electrical control and readout of single quantum states, the creation of decoherence-free subspaces to locally enhance coherence, and the integration of photonics to both enhance the efficiency of quantum emitters and entangle nearby quantum bits. Recently, rare-earth ions in oxide semiconductor hosts—including silicon-compatible materials—have emerged with encouraging properties as single quantum memories, with impressively long coherence times and single-shot photonic readout.

3.4 DEVELOP ENHANCED SPECTROSCOPIC AND MICROSCOPIC TECHNIQUES WITH NONCLASSICAL LIGHT

3.4.1 Quantum Light Spectroscopy and Entangled Photon Sources

As chemists push forward in their quest for deeper knowledge of the important processes and mechanisms of molecular interactions, great pressure is placed upon the nature of the measurements used to probe such processes. Indeed, the measurements themselves are also physical processes, and the accuracy to which these measurements may be performed to help better understand chemical processes is governed by the laws of physics. At the molecular level and for small systems, the processes involved in the measurements are governed by the laws of quantum mechanics that theoretically place limits on the accuracy to which the measurements can be performed. The intrinsic uncertainty of the results of the measurement of complementary observables is derived from the Heisenberg uncertainty relationship. Other quantum constraints combined with these uncertainty relations impose limits on how accurately we can measure quantities with the given physical resources of a measurement tool.

Nonlinear Entanglement Spectroscopy

New high-resolution, noninvasive spectroscopic and imaging methods for chemical processes have been developed over the past two decades. This success has allowed scientists to probe important chemical processes at the sub-femtosecond and sub-micron levels, which was not previously possible. However, better detection and imaging methods are still needed at the nanoscale at very low powers. Several impressive magnetic and optical techniques have been developed for this purpose. In the optical regime, both linear and nonlinear optical methods (see Figure 3-15) have been well illustrated in the literature. For example, nonlinear optical methods, such as two-photon absorption (TPA; and emission), have enjoyed great success in probing (at a very sensitive level) organic

Image
FIGURE 3-15 Schematic diagram of a nonlinear optical experiment with entangled photons. Here, a pump photon of frequency ωp is down-converted into two photons with frequencies ω1 and ω2 and directed onto the sample (bacterial reaction center).
SOURCE: Schlawin et al. 2013.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

and inorganic molecular systems as well as biological cellular (tissue) processes (Varnavski and Goodson 2020). The quadratic intensity dependence of TPA methods allows for very strong focusing and localization of the optical field, giving a much better effective resolution than their linear optical counterpart. The use of nonclassical sources of light such as squeezed states as well as entangled photons may provide an opportunity to exceed these quantum-imposed limits.

Great enthusiasm exists for exploring quantum information in molecular systems. Constructing a molecular system with the appropriate electronic states as well as inter- and intramolecular interactions for a desired quantum effect is a beneficial goal. This requires substantial advancement in both the synthesis and electronic structure of the proposed molecular systems. However, and perhaps of equal importance, this goal also requires advanced measurements at a very sensitive level for the desired detection of the spectroscopic or microscopic signal. New measurements are now being developed to probe the molecular properties of molecules at an enhanced level, which may one day provide new insights into the molecular design of systems for QIS applications as well as provide a more fundamental understanding of the nature of electronic interactions and mechanisms. In describing quantum phenomena in molecular systems, one usually is seeking behavior that lies outside the norm of classical behavior. For example, one is interested in the processes of coherence and decoherence in optical spectroscopy. These processes are being further defined and developed continually by time-resolved and nonlinear spectroscopic techniques. Chemists are developing the instrumentation and theory for new frontiers in this direction. Molecular systems typically suffer from a loss of coherence (amplitude and phase) at room temperature on ultrafast timescales (femtoseconds). A difficulty in measuring quantum states in molecular systems stems from the fast decoherence or mixing with a homogeneous broadening ensemble environment that is typical in solutions or solids. The development of new experimental techniques to provide new information through nonlinear signals, ultrafast multidimensional systems, and entangled photon interactions could enable improvement in the measurement of quantum-defined processes in molecular systems.

3.4.2 Quantum Light in Molecular Spectroscopy

The use of nonclassical states of light for spectroscopy may provide new approaches toward understanding the physical properties of molecules beyond quantum coherences to include entanglement. There are several ways that one may generate a quantum light source. Photon entanglement occurs between two beams of light when the quantum state of each field cannot be described in the individual parameter space of that field (Mukamel et al. 2020). While there are different parameters of light that can be entangled, the most commonly reported methods involve time and energy, polarization, and position and momentum. Along with the generation of squeezed states of light, these methods provide the bulk of the experimental techniques with nonclassical states of light. The interest in utilizing nonclassical states of light (e.g., entangled photons) stems from the potential advantages they may provide in comparison to classical laser excitation. It may be possible to take advantage of the high degree of temporal and spatial correlations in quantum light resulting from entangled pairs of photons (i.e., the signal and idler). These correlations can provide a sensing tool for chemists to detect potential molecular systems at very low levels of light, which is useful for investigations of chemical and biological systems. The pairs of photons are typically created by the process of spontaneous parametric down-conversion (SPDC) (Figure 3-16). These entangled pairs have shown different and low noise characteristics when compared to classical photon sources. This feature of entangled light can also be utilized in new linear and nonlinear spectroscopic approaches. All light fields have fluctuations in their amplitudes and phases, are subject to a stochastic indeterminacy beyond environmental influences, and involve a quantum indeterminacy. There has been a great deal of progress in producing light fields with fluctuations lower than the shot noise limit (SNL). One of these is amplitude squeezing, which may be produced by an interaction that preferentially removes the large amplitude fluctuations, resulting in a radiation field below the SNL. Both amplitude and photon number squeezing have attracted new applications in sensing and spectroscopy (Ispasoiu and Goodson 2000; Loudon and Knight 1987). Nonlinear four-wave mixing approaches have been carried out to produce bright squeezed states of light. Measurements are achievable with sensitivity beyond the standard classical limit and can be performed at low excitation fluxes for exciting photosensitive materials. With the advent of novel theoretical predictions and models, new measurements have been proposed that may suggest

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-16 Diagram of nonclassical light sources. On the left, spontaneous parametric down-conversion (SPDC) with a χ(2) bulk crystal showing the cone-like spatial arrangement of signal and idler photons for type-I/0 and -II SPDC and a periodically poled crystal (bottom left) with quasi-phase-matching properties. On the right, the phase-matching conditions and diagram of the four-wave mixing (FWM) process for creating squeezed light are shown, as well as an energy diagram of the involved beams.
SOURCE: Eshun et al. 2022.

an enhanced spectral resolution. Additionally, by utilizing the unique features of entangled photon pairs, one may also pursue the possibility of targeted entangled or coherent control of photochemical processes (Eshun et al. 2022). In the context of imaging (in biological systems), the highly correlated light provides the opportunity to probe chemical and biological systems at extremely low photon flux, avoiding the typically detrimental light toxicity issue. Entangled photon pairs provide a very powerful way to evaluate fundamental quantum mechanical principles such as Bell’s inequalities or in Hong–Ou–Mandel (HOM) photon correlation experiments (Schlawin et al. 2013).

Other intriguing properties of entangled photon spectroscopies include a non-Fourier-related spectral and temporal resolution. The spectral resolution is determined by the linewidth of the down-converted pump, whereas the temporal resolution is determined by the spectral width of the non-degenerate, down-converted photon pairs. This is because only the two entangled photons in a pair will interact with each other in an entangled photon spectroscopy measurement, whereas in a classical laser, any two photons can interact. Along these same lines is the ability to measure one photon using another wavelength or to image an object without spatially detecting the photons hitting the target. These “ghost” measurement techniques rely on the quantum correlations between the two photons in the entangled pair to determine the property of one photon using the other. It is notable that the photons must still be interfered with or measured by coincidence counting circuits to determine these properties. Such techniques are promising because extremes such as measuring an X-ray photon using a visible light photon are possible, all while gaining the quantum advantages of the entangled photons.

As depicted in Figure 3-16, SPDC is the most common method for creating nonclassical, entangled light. In the case of SPDC, a strong pump beam of frequency ωp interacts with a χ(2) nonlinear crystal to create a pair of lower-energy photons (the signal ωs and idler ωi photon), whose energy adds up to that of the pump, ωs + ωi = ωp (Eshun et al. 2022). In addition to energy, momentum also must be conserved, which leads to “phase-matching” conditions defined by ks + ki = kp. The phase-matching conditions determine the spatial correlations, efficiency, bandwidth, and temporal constants of the generated entangled photon pair. The most common nonlinear crystals used for SPDC are bulk β-barium borate (BBO) and potassium dideuterium phosphate (KDP). Because of said phase-matching

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

conditions, bulk crystals usually have efficiencies on the order of 10−10 to 10−12. The phase-matching conditions of bulk BBO and KDP also dictate tens of nanometer bandwidths and hundreds of femtosecond correlation times. These parameters put constraints on potential measurements to be performed with entangled photons.

The efficiency, bandwidth, and temporal correlations of an entangled photon source can be improved by using periodic poling of ferroelectric materials to create quasi–phase matching. Materials made popular by telecom entangled photon experiments include potassium titanyl phosphate and lithium niobate. Switching the poling of the crystals on the micrometer scale allows for artificial optimization of the phase-matching conditions. Higher-strength nonlinear tensors can also be used for type 0 SPDC. This results in a dramatic increase of the SPDC efficiency, to as high as 10−6, while also allowing a broader tuning of the central wavelength and bandwidth (and therefore temporal correlations) of the SPDC. Octave-spanning bandwidths with ~10 fs correlation times have been achieved with μW fluxes that are visible by eye. The greater bandwidth is accomplished by the introduction of a chirp in the poling field, which allows more wavelengths to be phase matched while also retaining the benefits associated with longer crystals (Szoke et al. 2021). Quasi–phase matching can be achieved in bulk crystals as well as waveguided sources to provide accessibility in both on-chip and free-space measurements.

A unique aspect of chemistry that is different from quantum information is the ability to create entangled photons over a broad range of central wavelengths. New proposals to use lithium tantalate have been published, offering the ability to pump SPDC down to ~350 nm where lithium niobate begins to suffer from self-absorption issues (Figure 3-17). Of course, even with lithium tantalate, the down-converted entangled photons are still in the 700 nm range, which is too high for most molecular dyes. Recent work has shown the creation of deep ultraviolet entangled photons using LaBGeO5 (LBGO). Entangled photons can also be created in the IR to terahertz regions using lithium niobate (<5 μm) and up to the terahertz range using orientationally patterned GaAs. Future work involves the use of sub-wavelength structures to further enhance the entangled photon generation rate (Price 2022).

As entangled light sources continue to be developed, attention has increased recently toward carrying out spectroscopy with the available entangled photons by current methods. The interest in utilizing entangled photons as spectroscopic tools first emerged in 1990, when scientists theoretically predicted a linear, rather than quadratic, scaling of the TPA rate with the pump photon intensity. One may rationalize this effect by comparing the interaction of the three-level atom with ordinary coherent light and with SPDC entangled light. If the atom interacts with coherent light tuned to the two-photon resonance, the two-photon transition rate may be estimated as the rate of excitation from the (virtual or real) intermediate state multiplied by the probability that the atom is in the intermediate state. At low intensity, both these quantities are proportional to the intensity, and one obtains the usual quadratic dependence of the two-photon rate on intensity. With the SPDC light, on the other hand, the excitation is accomplished in a single step: one photon of the pair promotes the atom to the virtual intermediate state, while its twin immediately (in a time less than the virtual-state lifetime) completes the two-photon transition. One may estimate the two-photon transition rate as the probability of excitation by a photon pair multiplied by the rate of arrival of photon pairs (proportional to intensity). The rate should be linear in intensity. The experimental verification of this effect in atomic and molecular systems established entangled photons as unique light sources for nonlinear spectroscopy with low photon fluxes (Figure 3-18). Measurements in atomic and molecular systems have been carried out for this form of entangled two-photon spectroscopy. Measurements of the entangled two-photon fluorescence with as few as 107 photons/sec have been carried out with this method. This gives rise to the possibility of doing spectroscopy with a small number of photons. The exact manner in which the entangled photons interact with particular molecular states and its impact on our understanding of the fundamental photochemical mechanisms could have important ramifications in our ability to excite and possibly control photochemical processes. Theoretical proposals for entangled virtual-state spectroscopy or entanglement-induced two-photon transparency have pointed out the highly unusual bandwidth properties of entangled photons and how this may be used further in understanding and possibly controlling chemical processes (Burdick, Schatz, and Goodson 2021; Eshun et al. 2022; Georgiades et al. 1995; Giri and Schatz 2022; Guzman et al. 2010; Harpham et al. 2009; Kang et al. 2020; Lee and Goodson 2006; Li et al. 2020; Parzuchowski et al. 2021; Raymer et al. 2013; Saleh et al. 1998; Schlawin, Dorfman, and Mukamel 2018; Svidzinsky et al. 2021; Tabakaev et al. 2021, 2022; Upton et al. 2013; Varnavski, Pinsky, and Goodson 2017; Villabona-Monsalve et al. 2018; Ye and Mukamel 2020).

Schlawin, Dorfman, and Mukamel (2018) have been instrumental in making theoretical predictions involving entangled photon spectroscopy. Nonclassical intensity fluctuations can enhance nonlinear optical signals relative to

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-17 (a) Spontaneous parametric down-conversion emission bandwidth as a function of chirp parameter for lithium tantalate, defined as a total percentage deviation from the degenerate poling period. (b) Numerically calculated full width at half maximum (FWHM) bandwidth trend as a function of poling period chirp.
SOURCE: Szoke et al. 2021.

linear absorption. This enables nonlinear quantum spectroscopy of, for example, photosensitive biological samples at low light intensities. In addition to the signal’s scaling with intensity, the use of entangled photons may also provide a new approach to shaping and controlling excitation pathways in molecular aggregates in ways that cannot be achieved with shaped classical pulses. For example, in some cases the use of entangled photons may suppress background signals and enhance others. In one proposed experiment, time–energy entangled photons produced by SPDC are employed to calculate vibrational hyper-Raman (HR) signals of a conjugated organic chromophore. Compared with classical light, Schlawin, Dorfman, and Mukamel (2018) found that time–energy entanglement can provide selectivity of Liouville-space pathways and will suppress the broad electronic-Raman background arising from one-photon resonances of the intermediate states. They showed that one-photon resonances can significantly

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-18 The linearity of various entangled two-photon absorption (ETPA) processes as a function of power. (a) Two-photon transition rate in trapped cesium with entangled and coherent light. The transition rate for uncorrelated coherent excitation is reduced by a factor of 10 for comparison. (b) Power dependence of ETPA rate in a porphyrin dendrimer under three different entanglement times. (c) Power dependence of sum-frequency generation with entangled photons. (d) ETPA rate for RhB and ZnTPP in solvent. (e) ETPA-induced fluorescence rate for Rh6G in ethanol under different concentrations. (f) Resonantly enhanced sum-frequency generation with entangled photons.
SOURCE: Szoke et al. 2020.

enhance classical HR signals; however, the electronic-Raman background is also introduced. By pathway selectivity, time-energy entangled photons can suppress this background while retaining the intense and narrow HR peaks (Figure 3-19). Another advantage of entangled light for HR spectroscopy is that the signal scales linearly with the pump intensity rather than quadratically. At low field intensities, entangled photons give stronger signals than

Image
FIGURE 3-19 Schematic of hyper-Raman (HR) spectroscopy with entangled photons generated by spontaneous parametric down-conversion (SPDC).
SOURCE: Chen and Mukamel 2021.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

classical light, making this tool valuable for experiments using fragile biological samples (Dorfman, Schlawin, and Mukamel 2014; Roslyak, Marx, and Mukamel 2009; Schlawin, Dorfman, and Mukamel 2018).

Other proposals from Dorfman, Schlawin, and Mukamel (2014) include using interferometric signals for pathway selectivity; applying multidimensional spectroscopy to suppress uncorrelated background signals; and employing interferometric TPA techniques with a triphoton entangled state, which will give stronger spectrally dispersed photon-counting signals than a typical biphoton state. The fundamental idea behind these proposed techniques is two-photon interferometry, which is best described with a HOM interferometer (Hong, Ou, and Mandel 1987). In a HOM interferometer, the entangled photon pairs are separated, and a time delay is introduced between them via differences in their propagation paths. Next, the photons meet and interfere at a 50:50 beam splitter, where single-photon detectors measure whether the photons emerge from the same or opposite sides of the beam splitter. The indistinguishability of the photon pair leads to destructive interference of the indistinguishable paths and a subsequent drop in the coincidence counts known as the “HOM dip” (Figure 3-20). Factors that distinguish the photon pair will affect the shape of the “HOM dip.” The dip reflects any matter interactions that the entangled photons encountered, which is an important feature for spectroscopic measurements. The increased sensitivity of quantum interferometers can improve the accuracy of measurements of material linear and nonlinear susceptibilities. This technique was first demonstrated in solid-state crystals and nanostructures with relatively narrow absorption lines. Earlier reports described coherent dynamics and dephasing processes in these systems. More recently, the HOM experiment was utilized to extract a dephasing time in an organic of ~102 fs upon coherent excitation and quantum interference with a path of entangled photons in the HOM interferometer. With theoretical modeling of the coincidence rate of the HOM dip by Eshun and colleagues (2021), the experiment and theory were analyzed and showed good agreement. The proof-of-concept report gives encouragement for further measurements of nonlinear signals with this quantum interferometry tool.

For entangled two-photon absorption (ETPA), and other processes where the photon entanglement is destroyed as a result of the excitation process, uncertainties remain about which proper final-state density to use in Fermi’s golden rule for determining the rate of absorption. For classical (unentangled) TPA, this state density is determined by the states of the molecule associated with the doubly excited state; normally, this is dominated by the density

Image
FIGURE 3-20 Hong-Ou-Mandel (HOM) pathways, spontaneous parametric down-conversion (SPDC) generation, and HOM interferometer setup. (a) Schematic of possible HOM interferometer pathways. (b) Femtosecond laser and SPDC generation with type-II SPDC β-barium borate (BBO) crystal. (c) Scheme of HOM interferometer. (d) HOM dip measured without sample.
SOURCE: Eshun et al. 2021.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

of phonons associated with this state, as this leads to the known rapid dephasing (in ~10 fs) after photoexcitation. However, for entangled photon excitation, Kang and colleagues (2020) proposed that entanglement fidelity can constrain the state density, such that the radiative lifetime of the doubly excited state rather than phonon dephasing determines this density. This result, which previously had been used in atomic physics modeling (Fei et al. 1997; Kojima and Nguyen 2004), leads to a much narrower lineshape for ETPA, and much larger cross sections are obtained from electronic structure calculations that are similar to what has been found in experiments (Kang et al. 2020). However, this assumes that fidelity of the entanglement constrains coupling to phonons, which is an assumption of unknown validity. Further work is needed to relate this result to known electronic and electron-vibrational entanglement properties of molecules (McKemmish et al. 2011; Plasser 2016).

There have been reports of utilizing nonclassical states of light in spectroscopy and microscopy in the linear optical regime. The use of quantum optical approaches in the linear regime could be of great interest to chemists in providing sensitive detection schemes for possible remote detection of chemical and biological systems. For example, there have been reports of quantum imaging with undetected photons. In this quantum interference effect, the photons that pass through the imaged object are never detected. The entangled photons that never interacted with the object provide the image. Here, it is not necessary to detect the photons that illuminate the object at all. This could be very useful for chemical and biological samples, where the material could lead to very opaque and high scattering of the illuminating light (Lemos et al. 2014). There have been other uses of quantum interferometers for linear optical effects in the IR regions. For example, reports have shown that by detecting the signal photons one can acquire the amplitude or phase information of the idler photons. Research has shown that the measured transmission spectrum of a polymer sample is in good agreement with a conventionally measured reference spectrum by detecting the correlated visible light of the SPDC process. Again, this approach could be useful for chemists interested in collecting IR spectral signatures of molecules using a visible laser light source and a silicon-based detector (Lindner et al. 2020; Mukai, Okamoto, Takeuchi 2022). This approach has even been extended to the terahertz region, where terahertz photons interact with a sample in free space and information about the sample thickness is obtained by the detection of visible photons. The ability to do terahertz spectroscopy with detection of visible photons would impact the investigation of thick opaque materials and samples (Kutas et al. 2020).

The use of entangled photons for coherent control has been reported as well by Gu and colleagues (2021). Recent theoretical work revealed that entangled light may provide a coherent control scheme for nuclear wave packets in a one-photon dark excited state of a molecule. This is demonstrated by nonadiabatic conical intersection wave packet dynamics for the transcis photoisomerization of azobenzene (Figure 3-21) (Casacio et al. 2021). This control leads to a substantial difference in the transient coherences during the passage through the intersection (the transition-state structure of the photochemistry), which is proposed to be detected by a stimulated X-ray Raman signal. Additionally, it was suggested that the photoisomerization yield is noticeably affected by modulating the photon entanglement time. The essential role of energy–time entanglement in the control is clearly demonstrated by contrasting the quantum light–excited wave packets to the wave packets created by classical light. Varying the bandwidth alone for both classical nonchirped and chirped pulses leads to minor differences in the two-photon excitation process. These results may provide a strategy for coherent quantum light control of the photoexcitation of electronic dark states of molecules (Gu et al. 2021).

3.4.3 Use of Quantum Light in Microscopy

One of the most recent applications of nonclassical states of light in chemistry involves microscopy, which is essential in broad areas of science from physics to biology. Using nonclassical states of light, chemists can open possibilities of realizing low-intensity microscopy at intensity levels not achievable with classical resources. For example, using quantum states of light for illumination, precise phase and linear absorption measurements in a microscopic format have been previously reported, and precision beyond the standard quantum limit has been achieved in a microscope (Casacio et al. 2021). These experiments also will be useful in overcoming photodamage, which is important for biological systems. Through the use of quantum correlations in nonclassical light, Casacio and colleagues (2021) demonstrated a highly selective coherent Raman microscope. They observed a 35 percent improved signal-to-noise ratio (SNR) compared to conventional classical light microscopy. This improvement

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-21 Azobenzene trans–cis isomerization initiated by entangled two-photon absorption. The combined energy of the two photons matches the S0/S1 transition at the trans geometry. The first reactive nuclear coordinate, q1, is the C1–N1–N2–C2 dihedral angle; the second coordinate, q2, is the symmetric bending of the C1–N1–N2 and N1–N2–C2 angles.
SOURCE: Gu et al. 2021.

translates to a 14 percent enhancement in concentration sensitivity, which reduces phototoxicity levels and allows for increased resolution of otherwise unobservable biological structures. The Raman microscope shows the vibrational spectra of biomolecules. Furthermore, quantum light eases the photon budget because it increases the SNR, as Raman scattering can be observed with even less than a single photon within the measurement time frame. Quantum-enhanced images of samples of both dry polystyrene beads and living Saccharomyces cerevisiae yeast were obtained (Figure 3-22) with 23 to 35 percent enhanced SNR, leading to increased image contrast. As such Casacio and colleagues (2021) were able to view features that were uncovered below the SNL.

Another approach toward the use of entangled photons in molecular and biological microscopy involves TPA (Eshun et al. 2022; Mukamel et al. 2020; Varnavski and Goodson 2020; Varnavski et al. 2022). Here, the microscopic image created by the fluorescence selectively excited by the process of the entangled TPA was reported. Entangled two-photon microscopy offers nonlinear imaging capabilities at an unprecedented low excitation intensity, 107, which is six orders of magnitude lower than the excitation level for the classical two-photon image. The nonmonotonic dependence of the image on the femtosecond delay between the components of the entangled photon pair is demonstrated. This delay dependence is a result of specific quantum interference effects associated with the entanglement, and this is not observable with classical excitation light. Additionally, fluorescent images of breast cancer cells were captured using entangled TPA in a scanning microscope (Figure 3-23; Varnavski et al. 2022). These images were generated at an excitation intensity orders of magnitude lower than the conditions necessary for classical two-photon microscopy. Quantum-enhanced entangled two-photon microscopy has shown cancer cell imaging capabilities at an unprecedented low excitation intensity of ~3.6 × 107 photons/sec, which is one million times lower than the excitation level for the classical two-photon fluorescence image obtained in the same microscope. The entangled two-photon microscope images resolving specific features of breast cancer cells in different stages of mitosis have been obtained. Varnavski and colleagues (2022) also illustrated the impact of

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-22 (a) Squeezed light normalized to shot noises, (b) stimulated Raman signal, and (c) signal-to-noise ratio (SNR) for a polystyrene bead. (d and e) Images of polystyrene beads and a live yeast cell taken with the quantum coherent microscope.
SOURCE: Casacio et al. 2021.
Image
FIGURE 3-23 Entangled two-photon light microscopy images of MCF7 cancer cells stained with the dye Hoechst 34580. (a and c) Cells with different numbers of nuclei. (b) Image of a colony of MCF7 cells. (d) Image of the large cell cluster. Insets: respective cell images created in a fluorescence microscope with the excitation by classical laser light at 405 nm. Spatial scale bar: 20 μm.
SOURCE: Varnavski et al. 2022.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

the low light excitation with entangled photons on the photodamage of the cancer cells. This is encouraging in the development of methods with extremely low light probe intensity with entangled two-photon microscopy critical to minimizing photobleaching during repetitive imaging and damage to cells in live-cell applications.

3.4.4 Measurement of Coherence in Molecular Systems

The measurement of coherence in chemical systems has a long history. The development of better methods to resolve more information about the molecular system is a significant part of modern physical chemistry. The instrumentation utilized in measuring the relatively fast dephasing times in molecules was significantly improved with the advent of the femtosecond laser, the use of which is now rather commonplace. As an example, one can utilize short laser pulses to excite ladders of states in phase, thereby making a superposition that can be detected using 2D spectroscopy. In the past decade, the number of research groups with the experience and equipment to carry out electronic and vibrational 2D spectroscopy has significantly increased. In this approach, a cross-peak in the 2D map labels the excited and detected transitions (Figure 3-24), while oscillations in the cross-peaks as a function of pump–probe waiting time reveal coherences involving those transitions occurring together. However, over time the oscillations damp away as a function of waiting time (Figure 3-24c). This spectral characteristic is a consequence of dephasing, giving a lower bound for the decoherence time of the quantum superposition. Figure 3-24a shows an example of electronic coherence for a semiconductor “nanoplatelet” colloid dispersed in

Image
FIGURE 3-24 (a) Absorption spectra of CdSe nanoplatelets (black line) showing the heavy-hole (HX) and light-hole (LX) exciton transitions. The spectrum of the laser pulses used in the two-dimensional (2D) spectroscopy experiments is shaded orange. (b) 2D electronic spectrum recorded at a pump–probe delay time of 52 fs. (c) Amplitude oscillations in the lower cross-peak of the rephasing 2D spectrum for a CdSe/CdZnS nanoplatelet (the real part with population relaxation subtracted) as a function of the waiting time. (d) A contour map of broadband pump–probe data for cresyl violet solution showing the oscillatory modulation on top of the ground- and excited-state population dynamics. (e) Fourier-filtered pump–probe data revealing coherent oscillations of the strong Franck–Condon active modes.
SOURCE: Scholes et al. 2017.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

an ambient-temperature solution. Here, the broadband femtosecond pulses (red curve) overlap with the first two exciton states, which are the heavy-hole (HX) and light-hole (LX) exciton. As demonstrated in Figure 3-24b, the amplitudes of the cross-peaks, HX-LX and LX-HX, in the 2D signal map oscillate as a function of excitation-detection time delay, showing that the amplitudes of HX and LX bands indeed are correlated until the superposition dephases. The dephasing of this ensemble happens with a time constant of 13 fs, which is aligned with previous observations for similar systems (Figure 3-24c). Electronic coherences at ambient temperature typically decohere with a time constant of ≤100 fs.

3.4.5 Coherence and Biological Function Is the Future

In the past 10 years, electronic and vibrational coherence experiments have improved significantly with the use of time-resolved nonlinear spectroscopic techniques. Mukamel and colleagues (2020) have been instrumental in the theory and prediction of spectroscopic signals under this context. Chemists have investigated the electronic coherences in organic and biological systems in impressive detail. According to Scholes and colleagues (2017), it might be appropriate for the field to move beyond measuring the coherence lifetimes. The use of multidimensional spectroscopic methods to probe other parameters related to the excited-state properties could provide new insights into the connection between biological function and chemical structure. For example, time-resolved ultrafast optical probes of chiral dynamics may provide a new window to explore how interactions with structured environments drive electronic dynamics. Incorporating optical activity into time-resolved spectroscopies has proven challenging owing to the small signal and large achiral background. Higgins, Allodi, and colleagues (2021) demonstrate that 2D electronic spectroscopy (ES) can be adapted to detect chiral signals and that these signals reveal how excitations delocalize and contract following excitation. They dynamically probe the evolution of the chiral electronic structure in the light-harvesting complex 2 of purple bacteria (Figure 3-25) following photoexcitation by creating a

Image
FIGURE 3-25 Structure and model of light-harvesting complex 2 of the purple bacteria reaction center.
SOURCE: Higgins, Allodi, et al. 2021.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

chiral 2D mapping. The dynamics of the chiral 2D signal directly report on changes in the degree of delocalization of the excitonic states following photoexcitation. The mechanism of energy transfer in this system may enhance transfer probability owing to the coherent coupling among chromophores while suppressing fluorescence that arises from populating delocalized states. This generally applicable spectroscopy will provide an incisive tool to probe ultrafast transient molecular fluctuations that are obscured in non-chiral experiments. Although coherence has been shown to yield transformative ways for improving function, advances have been confined to pristine matter and coherence was considered fragile. However, recent evidence of coherence in chemical and biological systems suggests that the phenomena are robust and can survive in the face of disorder and noise. Chemists are still assessing the variability of this process and the possibility that coherence experiments and mechanisms can be used effectively in more complex chemical systems. A key missing point is the role of coherence as a design element in realizing function.

Most recently, quantum coherences observed as time-dependent beats in ultrafast spectroscopic experiments arise when light–matter interactions prepare systems in superpositions of states with differing energy (e.g., redox states) and fixed phases across the ensemble. These measurements reveal important variations when compared to previous reports, suggesting that redox conditions tune vibronic coupling in the Fenna–Matthews–Olson (FMO) pigment–protein in green sulfur bacteria; this raises the question of whether redox conditions may also affect the long-lived (>100 fs) quantum coherences observed in this very important complex. Higgins, Lloyd, and colleagues (2021) performed ultrafast 2DES on the FMO complex under both oxidizing and reducing conditions (Figure 3-26). They observed that many excited-state coherences are exclusively present in reducing conditions and are absent or attenuated in oxidizing conditions. Reducing conditions mimic the natural conditions of the complex more closely. Furthermore, the presence of these coherences correlates with the vibronic coupling that produces faster, more efficient energy transfer through the complex under reducing conditions. The growth of coherences across the waiting time and the number of beat frequencies across hundreds of wavenumbers in the power spectra suggest that the beats are excited-state coherences with a mostly vibrational character whose phase relationship is maintained through the energy transfer process. These results suggest that excitonic energy transfer proceeds through a coherent mechanism in this complex and that the coherences may provide a tool to disentangle coherent relaxation from energy transfer driven by stochastic environmental fluctuations.

As this field shifts its focus from confirming the existence of coherence to exploring the connection between coherence and function, new methods are being extended to provide new insights. As expected, this area of investigation will require extensive feedback between theory and experiment, synthesis and measurement, and the development of systematic methods to quantify the influence of coherence in specific processes or devices. Exploring function requires controlled perturbation, and establishing this essential methodology requires new control mechanisms and clear assessment tools. For instance, new and improved experimental techniques will enhance capabilities such as measuring the delocalization of wave functions or measuring the collapse of delocalized states. In this regard, attosecond laser sources have enhanced the ability to study coherent electronic motion, such as charge migration and quantum interference between electrons, as well as electron-nuclear wave packet motion. Similarly, advances in time-resolved X-ray spectroscopies open up new probes of electronic structure by delineating dynamics into element-specific contributions. Inspired by coherent multiple scattering and molecular transport junctions, the reactivity of metal centers might be directed or redox chemistry tuned by interference effects to modify the electron density at an active site of a catalyst. In coherent backscattering, which provides scattering that is twice as bright as diffuse scattering, substantial gains are possible when robust coherence phenomena are exploited. Such gains warrant future research into coherence as a potential force for enhancing function.

3.4.6 High-Finesse Cavities and Nanophotonics for Molecular Qubit Systems

In addition to understanding and designing molecular systems with favorable optical properties, a different and complementary approach is to engineer optical control by modifying the molecule–light coupling, by changing the vacuum field of the light itself. One way to accomplish this is by surrounding a molecule with an optical cavity (Figure 3-27). Light resonant with the cavity recirculates and repeatedly interacts with matter located in the cavity modes, leading to enhanced light–matter coupling. In addition, the cavity can also modify the decay rate of optically excited molecular states by modifying the density of states of the electromagnetic field.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-26 Redox condition affects excited-state behavior in the wild-type Fenna–Matthews–Olson (FMO) complex. (a) Structure of the FMO complex and the eight bacteriochlorophyll sites held by the protein scaffold (Protein Data Bank ID code 3ENI). Shown in red are the two cysteine residues, C49 and C353, that are known to steer and quench excitations in oxidizing conditions and tune vibronic coupling for enhanced energy transfer in reducing conditions. (b) Linear absorption spectra of the wild-type oxidized (blue) and wild-type reduced (red) FMO complex at 77 K. Shown in gray is the laser spectrum used. (c and d) Rephasing two-dimensional electronic spectra under oxidizing and reducing conditions at waiting time T = 40 fs. Differences in the lower-diagonal cross-peaks between experiments indicate faster, more efficient energy transfer when the complex is reduced.
SOURCE: Higgins, Lloyd, et al. 2021.

The applications in quantum science from surrounding molecules with cavities are many and include building quantum networks (Duan et al. 2001; Kalb et al. 2017), creating photon-mediated interactions between quantum emitters (Evans et al. 2018), and building single-photon sources. Here, we discuss a few examples with relevance to optical control of molecules.

First, with regard to the ZPL problem encountered in solid-state qubits, cavities can enhance emission into the ZPL via the Purcell effect, as has been proposed and experimentally demonstrated (Barclay et al. 2011; Riedel et al. 2017). In the Purcell effect, the decay rate of a transition resonant with an optical cavity is strongly enhanced (Purcell, Torrey, and Pound 1946). This leads to preferential emission into a specific state. State-of-the-art experiments with NV centers in a microcavity have shown enhancement of ZPL emission from 3 to 46 percent, with a

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-27 Illustration of optical microcavities that may one day allow chemists to control chemical reactions.
SOURCE: Du, Ribeiro, and Yuen-Zhou 2019.

viable path toward emission factors exceeding 90 percent (Riedel et al. 2017). Second, optical cavities can provide coherent interactions between matter spaced over large distances. For atomic systems, this has been proposed as a way to produce entanglement and teleportation of quantum information over long distances. The same principles apply to optically active molecular systems.

Although we have focused above on how optical cavities can be used to produce more favorable qubits, combining the QIS technique of optical cavities with molecules also opens a potential new way to control chemical reactions. In the presence of strong light–matter coupling, molecules within a cavity inherit photonic character and form so-called polaritons, which are quantum superpositions of light and matter. This has led to much interest recently in polariton chemistry (see below), where new possibilities abound. With dense highly absorbing molecular ensembles in microfluidic cavities and dye lasers, experiments already have seen polaritons affect chemical reactivity and optoelectronic properties (Ebbesen 2016; Ribeiro et al. 2018). New types of molecular polaritons with unique properties could be synthesized (Gu and Franco 2018), and novel catalysts might be found. For example, with molecules in cavities, it could be possible to control photochemistry remotely simply by tuning a “remote catalyst” in a distant cavity that is coupled to the system of interest via photons (Du, Ribeiro, and Yuen-Zhou 2019). The many new possibilities with molecular systems in optical cavities come with some technical and material challenges, including fabricating cavities with low loss (i.e., high reflectance and low scatter) and integrating cavities with molecular systems.

Analogous to the work with optical cavities described above, demonstrations of strong spin–photon coupling in resonant microwave devices is another area of significant exploration (Bonizzoni, Ghirri, and Affronte 2018; Eddins et al. 2014; Gimeno et al. 2020; Jenkins et al. 2013). All of these investigations have been in a regime far from the quantum limit, involving macroscopic numbers of spins and photons. Nevertheless, strong coupling effects have been reported for a number of different molecular systems and resonator types. This is an area ripe for further development in order to reach a point where entanglement between microwave photons and a small ensemble of molecular spins (or even a single molecular spin) is achieved.

Cavity Polaritons

When strongly coupled, material states and optical cavity modes transform into a single quantum entity—a polariton—with its own distinct and modifiable physics and chemistry, fundamentally altering the behavior of both matter and the electromagnetic field. Cavities may provide a nondestructive and versatile way to control energy, charge, and coherence transfer in molecular systems, which underlie reactivity, photochemistry, and catalysis. To

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

establish the fundamental scientific understanding of cavity-enabled molecular control, it is necessary to establish basic principles that can guide future chemical innovation. With these essential concepts identified, the modular nature of external cavity control will usher in the rapid technological development of cavity-enabled chemistry. By leveraging, for example, microfluidic and lab-on-a-chip platforms, it will be possible to carry out complex, high-value chemical transformations that are key to the production of pharmaceutical and industrial chemicals. This could indeed be another avenue for chemistry in QIS.

The use of cavities is already under way in chemical research, with two parallel but related efforts in the polaritonic chemistry community. One thrust has been to apply ultrafast laser spectroscopy methods to monitor the complex energy transfer and relaxation processes in molecular polaritons, both in the vibrational and electronic domains. While considerable basic science interest certainly exists, there are also many possibilities for applications in QIS using molecular polaritons as quantum resources, particularly in the case of quantum transduction. Vibrational polaritons are especially promising in this context because of the relatively long coherence times achievable using molecular vibrations coupled to resonant optical cavities. The use of ultrafast 2D optical spectroscopy, both in the visible and mid-IR spectral regions, has revealed nonlinear optical effects of potential use for optical switching, polaritonic energy ladders, and the exchange of energy between bright polariton states and nearly inactive dark states (Figure 3-28). There are clearly challenges that require advanced cavity designs, as suggested by a recent study showing considerable contributions to the measured two-dimensional infrared (2DIR) response from molecules that are neither polaritons nor dark states but instead comprise a non-polaritonic background (Duan et al. 2021). These fundamental dynamics studies are important because they directly probe the underlying physics of all the components present in a molecule-cavity construct.

The other major thrust in polaritonic chemistry is arguably the most promising avenue for future investigation and discovery. Following the first observation of a cavity-controlled photochemical reaction by Laboratoire des Nanostructures at the University of Strasbourg, there are now several examples of chemical reactions whose kinetics and equilibria can be altered through coupling to a resonant optical cavity (George et al. 2015). An important recent finding showed that the selectivity and rate of bond cleavages in an alkyl silane compound can be modified within a cavity tuned to specific resonances of the reacting species (Nagarajan, Thomas, and Ebbesen 2021). This observation implies that cavities can exert both thermodynamic control (e.g., altering the equilibrium constant of a reaction) as well

Image
FIGURE 3-28 Two-dimensional infrared spectra of vibrational polaritons.
SOURCE: Duan et al. 2021.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

as kinetic control by selectively lowering the barrier for one reaction channel relative to another. Perhaps an exciting recent development is a report that shows that a urethane addition reaction can be controlled by strong coupling to a cavity (Ahn, Herrera, and Simpkins 2022). Though no more fundamental chemically than the alkyl silane example, this latter work is the first by a research group that is completely independent of the University of Strasbourg team. Indeed, challenges in reproducing some cavity results have cast some doubt on the robustness of cavity control, but the recent urethane addition reaction provides ample evidence that cavity control is possible in other laboratories. Theoretically, scientists such as Gu and Mukamel (2021) are working to describe this process in detail.

Perhaps the key outstanding challenge that the field must confront is the lack of a clear set of guiding principles that underlie the mechanism(s) of cavity control. Whereas the physics of strong coupling between molecules and cavity modes are understood, no consensus exists regarding the reasons why creating delocalized states with energies slightly above and below the uncoupled energies would cause reaction rate constants to increase or decrease to the degree reported in the literature. Statistical theories, such as transition-state theory, that routinely apply broadly in chemistry seem to be completely inconsistent with polaritonic effects on reactivity. Dynamical approaches based on quantum electrodynamic density functional theory are feasible but are extremely demanding and are unlikely to be helpful for the broader chemistry community. A set of experimental results on tractable chemically reactive systems that can be studied to provide inspiration and inputs to theoretical analyses will deepen this research area. Ideally, these reactions will be elementary and will not rely on complicated multistep mechanisms. Reactions that are phototriggered will enable direct dynamical investigation rather than more coarse-grained kinetics measurements. In some sense, polaritonic chemistry, especially when coupling vibrational transitions to cavity modes, is the fulfillment of a dream that began with the advent of lasers in the 1960s that we would be able to do chemistry precisely using mode selectivity. A promising aspect of the present situation is that it is highly likely that basic principles will be established quickly—particularly with enhanced funding support—paving the way for accelerated chemical technologies that will offer entirely new ways to conceive of and carry out chemistry.

Another area of research that has received primarily theoretical attention in the chemistry community is the use of quantum light in the context of polaritonic effects in cavities. Gu and Mukamel (2020) have shown that nonlinear quantum light spectroscopy can be employed for probing two-photon excitations in polaritonic systems. Theoretical investigations have identified the combined signatures of a cavity photon mode strongly coupled to molecules and entangled photons on collective bipolariton resonances. Initial studies focused on TPA signals with an entangled photon pair to a polaritonic system consisting of N two-level molecules strongly coupled to a single cavity photon mode. By comparing the TPA spectra with classical and quantum light, Gu and Mukamel (2020) have demonstrated that entangled photons can create two-photon excitations in polaritonic systems that are drastically different from the classical two-photon excitations. Entangled photons can reveal classically dark bipolariton states by modifying the quantum interference among transition pathways leading to TPA (Figure 3-29). These exciting predictions have yet to be examined experimentally and provide a novel opportunity for chemists working in QIS.

Image
FIGURE 3-29 Illustration of two-photon excitations to bipolariton states created by placing several molecules in an optical cavity excited by quantum light.
SOURCE: Gu and Mukamel 2020.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-30 Illustration of a cavity-manipulated singlet fission system that is mediated by polaritonic conical intersections for both one- and two-molecule systems.
SOURCE: Sun, Gelin, and Zhao 2022.

Other theoretical studies have shown that cavity-manipulated singlet fission (SF) is possible when it is mediated by polaritonic conical intersections for both one- and two-molecule systems (Figure 3-30). The population evolution of the critical state for SF (Triplet-triplet state) and the cavity photons were carefully examined in search of a high fission efficiency via cavity engineering. Several interesting mechanisms were uncovered, such as photon-assisted SF, system localization via a displaced photon state, and collective enhancement of the fission efficiency for the two-molecule system. It was also found that the system localization process in the two-molecule system differs substantially from that in the one-molecule system owing to the appearance of a novel central polaritonic conical intersection in the two-molecule system (Sun, Gelin, and Zhao 2022). The possibility that a cavity-controlled SF process can be switched on and off by controlling the average pumping photon number is another novel opportunity for chemists working in QIS.

3.5 DEVELOP AND EXPLOIT ALTERNATIVE APPROACHES TO SPIN POLARIZATION AND COHERENCE CONTROL

3.5.1 Chirality-Induced Spin Selectivity Effect

Recent experiments have shown that it is possible to generate high spin polarization via electron transport through chiral molecules, even at room temperature. This discovery of the so-called chirality-induced spin selectivity (CISS) effect (Naaman and Waldeck 2012; Ray et al. 1999), whereby spins aligned parallel or antiparallel to the electron transfer displacement vector are preferentially transmitted depending on the chirality of the molecular system, is of huge potential importance for the molecular QIS research endeavor. In particular, the coupling of orbital and spin angular momenta during directional electron transfer processes provides a novel means for manipulating spins over molecular length scales. A detailed discussion of the fundamental chemistry and physics governing the CISS effect can be found in Section 2.4. Nevertheless, it is clear that an area ripe for future investigation involves optimization and exploitation of the CISS effect with potential molecular QIS applications in mind. Such efforts will rely on nanosecond and femtosecond transient absorption spectroscopies in order to probe electron transfer dynamics (Harvey and Wasielewski 2021). In addition, the use (and further development) of time-resolved/transient EPR spectroscopy will be essential to elucidating the CISS-induced spin dynamics.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

3.5.2 Electric Field Control

Various efforts have been pursued that employ EPR detection to monitor the response to an external stimulus applied to a sample. For example, slow quantum relaxation of non-equilibrium spin populations has been studied via high-frequency EPR detection (Dressel et al. 2003), as well as over-barrier relaxation due to thermal avalanches triggered by acoustic pulses (Macià et al. 2008). Meanwhile, in situ application of large hydrostatic pressures within a high-field EPR instrument enables fundamental insights into the nature of the microscopic interactions governing the spin physics in molecular materials (Thirunavukkuarasu et al. 2015), while also demonstrating structurally induced switching effects (Prescimone et al. 2012). More recently, several studies have demonstrated magnetoelectric coupling through EPR measurements carried out under static (or quasi-static) electric fields (Boudalis, Robert, and Turek 2018; Fittipaldi et al. 2019; van Slageren 2019). These investigations demonstrated an important milestone on the pathway to molecular QIS by enhancing prospects for local control at the nanometer scale, along with the obvious advantages of the low power consumption needed for the manipulation of quantum systems with electric fields. The first such investigation involved an antiferromagnetic triangular iron-oxo cluster. In this case, competing antiferromagnetic Heisenberg and Dzyaloshinskii-Moriya interactions give rise to spin chirality and the observed magnetoelectric coupling observed as an increase in EPR intensity under the application of a direct current electric field (Figure 3-31; Boudalis, Robert, and Turek 2018). More robust results were obtained subsequently through alternating current electric field modulation of exchange interactions in helical metal-radical chains (Fittipaldi et al. 2019).

Very recently, an entirely new form of coherent pump-probe EPR spectroscopy has been developed by Liu and colleagues (2019) to study spin–electric coupling in molecular qubit candidates. The method is analogous to the pulsed ELDOR and DEER techniques, except that a pump electric field pulse is employed in order to modulate the coherent spin dynamics via the magnetoelectric effect. As highlighted in Section 2.3, perhaps the most exciting

Image
FIGURE 3-31 An electric field applied on the antiferromagnetic spin-chiral complex [Fe3O(O2CPh)6(py)3]ClO4·py couples to the spin of its ground state and modifies its electron spin resonance spectrum. This magnetoelectric coupling can provide electric control of the spin states of molecular nanomagnets.
SOURCE: Boudalis, Robert, and Turek 2018.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

example involves the application to clock transitions associated with a Ho(III) molecular spin qubit (Liu, Mrozek, et al. 2021). Crystals contain two electrically polar molecules related by inversion. Importantly, the structural distortion that gives rise to the electric dipole moment is also responsible for the symmetry breaking that induces the clock transition (the avoided level crossing between the Zeeman levels corresponding to electronic spin-up and -down states)—that is, the clock transition frequency is directly coupled to the molecular dipole moment. In turn, this implies linear coupling of an electric field to the clock transition frequency. However, the frequency shift is opposite for the two inversion-related molecules. Consequently, the application of an electric field pulse during the first half of a Hahn-echo sequence results in an eventual phase-shifted refocusing of the two populations at angles ±φ relative to the magnetic pulses within the rotating frame, where the magnitude of φ scales linearly with the duration and amplitude of the electric field pulse. Thus, a periodic modulation and decay of the echo is observed in the in-phase echo signal, with no echo detected in the quadrature channel (due to cancelation of the ±φ shifted echoes from each subpopulation). The decay is caused by the inhomogeneous response of the ensemble to the electric field pulse. Since these inhomogeneities are static, the spin echo is completely recovered if the electric field pulse spans the second (refocusing) half of the Hahn-echo sequence (see Figure 3-32). It is then possible to perform a 2D experiment, after which the electric field pulse is applied only during the refocusing magnetic π pulse while scanning the frequency of the π pulse. Refocusing is only effective for one of the subpopulations

Image
FIGURE 3-32 (a) Structure of the two inversion-related polar [HoIII(W5O18)2]9− molecules and the experimental configuration in which the sample is subjected to a static applied magnetic field, B0, and pulsed electric fields, E (in addition to microwave electron paramagnetic resonance pulses). (b, c) Sequences of microwave magnetic (yellow) and electric (blue) field pulses and the resulting evolutions of the magnetizations associated with the oppositely polarized molecules, leading eventually to an echo (red). (d) Temporal evolution of the in-phase and quadrature spin-echo signals for the pulse sequence (b); dephasing occurs due to the inhomogeneous response to the electric field pulse. However, this dephasing is refocused if the E-field pulse is kept on for the full duration of the Hahn-echo sequence. (e) Applying the E-field pulse during the magnetic π pulse (c) breaks the symmetry between the two electrical polarizations.
SOURCE: Liu, Mrozek, et al. 2021.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

(i.e., half of the molecules) when the π-pulse frequency matches the shift caused by the electric field. Hence, the echo signal drops in intensity by a factor of two and shifts linearly both up and down in frequency in response to the magnitude of the electric field (see Figure 3-32). In this way, the electric field breaks the inversion symmetry inherent to the crystal, enabling selective excitation of the two inversion-related molecules. More generally, these studies clearly demonstrate coherent spin-electric control in molecular nanomagnets that exhibit long coherence times due to clock transitions (Shiddiq et al. 2016), paving the way toward local electrical control of molecular spin qubits at the nanometer scale with low dissipation (Ullah et al. 2022). These methods are now finding more general utility in the chemical study of magnetoelectric materials (Liu, Mrozek, et al. 2021).

3.6 EXTEND QUANTUM TELEPORTATION AMONG MOLECULAR QUBITS

Quantum communication is a rapidly growing area of QIS and engineering, with quantum network test beds appearing around the world. In contrast to classical communication, quantum communication takes place by sending or receiving information transmitted using the quantum states of a specific quantum system (e.g., a photon). Such “flying qubits” offer built-in security given that the (unwanted) observation of these states collapses the information. A quantum internet is being developed using entangled qubits to create a unique mode of information processing utilizing teleportation. In addition to security, a quantum network may enable quantum computing clusters, connecting multiple quantum computers to create more powerful machines for applications in science and engineering. Ultimately, one imagines a quantum ecosystem where sensors, communications (wiring), and computers exchange quantum information without exposure to the classical world. This new paradigm for communications requires the community to address a growing number of significant challenges, including the creation of quantum repeaters and robust quantum memories. These needs sit at the interface of fundamental physics and chemistry, representing exciting and important scientific research areas.

Today’s internet is driven by the existence of repeaters that circumvent our current limits in materials science. Digital optical pulses travel through optical fibers until their amplitude is diminished through scattering processes in the glass. At that point, the signal is read, amplified, and repeated downstream. This process is repeated until the signal reaches its destination. Given the laws of quantum physics, quantum repeaters will require a different operational mode given the inherent information collapse upon observation. To that end, the community currently envisions new concepts such as exploiting and combining different quantum degrees of freedom, as well as mixing photons, phonons, magnons, and spins in an effort to amplify one mode in lieu of another. Devising and constructing atomically engineered materials that enable quantum transduction at the single spin–photon and/or single phonon–photon level may be greatly advanced through molecular chemistry, where advanced theory and synthesis may be used to manipulate local bond strengths and spin–orbit interactions, integrate single ion memories, tune optical levels, and design integration pathways with optical communication pathways (fibers). In addition to a high degree of spatial control and tunable emission energies, there are opportunities to multiplex quantum signals within single structures given molecular length scales and existing organic molecular frameworks. The remarkable precision, placement, and tunability of molecular qubits and potential memories make molecular quantum systems powerful candidates to accelerate research in science and engineering (Baek et al. 2005).

Use Molecular Systems to Teleport Quantum Information Over Distances Greater than 1 μm with High Fidelity

Control of the coherence times of quantum states is currently one of the major challenges in QIS (Wasielewski et al. 2020). Spin-state decoherence can be accelerated greatly by interactions such as spin exchange, hyperfine coupling, spin–orbit coupling, and magnetic dipolar coupling. By controlling the structure and composition of molecular qubits, many of these decoherence sources can be mitigated and/or controlled (Krzyaniak et al. 2015; Yu et al. 2016). For example, purely organic molecular qubits have the advantages of weak spin–orbit coupling and well-defined electron–electron and electron–nuclear spin exchange interactions (Nelson et al. 2017). Moreover, ultrafast photochemical electron transfer within an organic donor-bridge-acceptor (D-B-A) molecule can produce a radical pair that can function as two entangled spin qubits (D•+–A•−), giving rise to an entangled two-spin singlet or triplet state (Krzyaniak et al. 2015; Yu et al. 2016).

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

As discussed in Chapter 2, this strategy has been used to achieve electron–spin state teleportation. This is essential to preserve quantum information across long distances (Pirandola et al. 2015) in an ensemble of covalent D-A-R molecules, in which it is possible to propagate the initially prepared spin state of a stable radical R to D•+ (Pirandola et al. 2015; Rugg et al. 2019). It is desirable that future research focuses on developing molecular systems that can achieve high-fidelity quantum teleportation of electronic states on length scales approaching 1 μm. Such systems could thus serve as quantum interconnects in larger-scale devices. In addition, as discussed in the following section, the development of chemical approaches that take advantage of entangled photons is also of importance for quantum teleportation/communication over ever greater distances.

3.7 DEVELOP MOLECULAR QUANTUM TRANSDUCTION SCHEMES THAT TAKE ADVANTAGE OF ENTANGLED PHOTONS AND ENTANGLED ELECTRON AND NUCLEAR SPINS

Quantum transduction refers to the transfer of quantum states between different forms of quantum subsystems to connect qubits, support quantum networks, or allow for quantum sensing. Current quantum transduction schemes range from electron spins, nuclear spins, bosons, acoustic modes, and microwaves to optical photons. To interface with larger quantum systems or classical input-output schemes, electrical or photon-based transduction is often assumed optimal, but all areas are being explored. Molecules offer a unique approach to quantum transduction because of the tunable interaction between nuclear, rotational, vibrational, and electronic degrees of freedom. Molecular transduction schemes offer more degrees of freedom than ion systems while being more energetically discrete than solid-state approaches. The instrumentation described in this chapter must be developed and combined to probe these potential forms of quantum transduction, as transduction naturally includes measuring the entanglement between multiple degrees of freedom and their modulation simultaneously.

The classic example of quantum transduction is the attempt to couple superconducting transmon qubits by converting electrical or microwave signals into optical photons (flying qubits). While RF and microwave hardware likely will remain central to efforts aimed at coherent control of electron and nuclear spins in molecules, the lack of sensitivity inherent to conventional electromagnetic detection modalities in these frequency bands prompts the exploration of alternative transduction schemes. Optical detection is known to provide single-spin sensitivity in the case of the NV center in diamond (Chen et al. 2017), and efforts focusing on optical sensing of molecular spin qubits were described earlier in Section 3.4 and later in Section 3.10. Another approach allows single-spin detection (Thiele et al. 2014) via electrical readout of a single-molecule transistor (Gehring, Thijssen, and van der Zant 2019; Jo et al. 2006; Pietsch et al. 2016) or chiral-induced spin selectivity and spin chemistry. Entanglement between a nuclear spin and a rotational mode has been measured, creating potential transduction routes through vibrations. Similar coupling to a surface substrate, perhaps through magnons, is also a feasible proposal (Godejohann et al. 2020).

For long-range coupling, there are opportunities to research how spin qubits interact with entangled photons or how photon-based “flying qubits” can be made from molecular qubits. Unique opportunities exist for creating quantum repeaters as optical circuit elements using the inherently spin-controlled optical and vibrational processes of molecules (Dolde et al. 2013). For example, photogenerated electron spins with classical or entangled photons are of great utility because they can be prepared in well-defined quantum states. Quantum transduction between electron and nuclear spins can then be used for storage elements; nuclear spins are highly localized and have ~1,000 times smaller gyromagnetic ratios than electrons, making them less sensitive to their environment and therefore enabling coherence lifetimes orders of magnitude longer than electron spins (Maurer et al. 2012; Saeedi et al. 2013). However, in order to optimize molecular qubit systems that take advantage of the properties of both electron and nuclear spins, three important questions need to be answered through instrumentation development:

  1. What are the limits of electron and nuclear spin coherence in designed molecular qubits?
  2. How do electron–nuclear spin–spin interactions influence coherence transfer between electron and nuclear spins?
  3. What combinations of microwave and optical photons are needed to optimize coherence transfer between multiple such subsystems?
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

For shorter-range coupling, chemistry enables the design of large molecular structures with atomic precision that can be used as local processing or sensing nodes. Hierarchical structures, such as metal–organic frameworks (Feng et al. 2020) and molecular crystals (Kothe et al. 2021), provide a platform for scaling up single-molecule qubits to functional multiqubit arrays. Molecular qubits can be connected through linkages that are only a few nanometers long, potentially leading to higher-density qubit arrays than those based on isolated atoms or defects. These scaffolds can facilitate the control of the entangling interactions between qubits, such as the spin exchange coupling between two electron spin qubits (Olshansky et al. 2019). Photoactive or switchable bridge molecules can provide a convenient platform for changing the sign and the magnitude of the magnetic exchange couplings between two spin qubits, enabling ultrafast state manipulation, changes in the polarization of individual qubits, and modulation of qubit entanglement. For example, using phenylene bridges, the strength of the coupling can be reduced by a factor of 50 through torsional distortions around the single C–C bonds joining the phenyl groups (Stasiw et al. 2015). Transduction through short-range dipolar-, electronic-, or vibrational-based routes within a dense array, outcoupled by a photon or electrical signal, can then allow for longer-range transduction.

In all of these cases, beyond synthesis needs, measuring quantum transduction requires new tool sets that can evaluate entanglement between multiple quantum subsystems simultaneously—for example, by combining an EPR measurement of spin coherence with a measurement of the emitted entangled optical photon. Control of a single or group of molecular qubits requires transduction, both locally and long range, and linkage to classical external feedback circuits. Methods to measure and realize molecular quantum transduction represent a significant scientific frontier.

3.8 ADVANCE QUANTUM SENSING TECHNIQUES TO FURTHER UNDERSTAND BIOLOGICAL SYSTEMS

As described earlier in this report, the area of quantum sensing is rapidly advancing with early proof-of-concept demonstrations that showcase its breadth. Recent advances in sensor technology suggest that quantum sensors have significant near-term potential. Chemistry is uniquely poised to transform this field. Sensors are inherently analyte specific and environmentally dependent. Developing designer quantum sensors is a challenge perfectly suited for synthetic chemistry. Deep literature exists on modifying chemical compounds for specific environmental compatibility. As an illustrative example, a key question for quantum sensing in biological systems relates to monitoring catalytic turnover in enzymes. Moving from imaging an ensemble of molecules to a single molecule or even hundreds would be transformative for our understanding of catalysis. Since the key steps in a catalytic cycle are very fast, statistically only a small number of molecules in an ensemble are in the critical state at the correct time. As such, rapid measurements that enable sub-ensemble size scales could elucidate the catalytic mechanism for important systems such as the oxygen-evolving complex at the heart of photosystem II. How would a chemical approach be suited to this? Initially, the target would be selected—for instance, readout of the spin state of the dangling Mn ion with light. We would need to use a molecular color center to execute the measurement with an emission frequency that was biologically compatible. Moving toward the lower energy frequency region of the electromagnetic spectrum is generally less biologically damaging and has better depth penetration. The first step would be to design a system suited to sense spin in this regime (Figure 3-33). Subsequently, we would need to introduce a molecule into a biological environment. The molecule would need to be small and biologically compatible, and ideally contain a chemical tether that would bind to the appropriate substrate. Tethering chemicals to biological systems is an old field, with DEER spin labels and fluorescence resonance energy transfer labels as canonical examples. For quantum sensing, there are additional design parameters beyond conventional sensing, coherence time, and temperature dependence (Figure 3-34). The tunability of molecules is very useful for designing these parameters.

3.8.1 Investigating Quantum Effects in Biology with Quantum Tools

Interaction between quantum physicists and biologists can be traced back to the beginning of the quantum era itself. However, as described earlier in Section 3.4, as the sensitivity and detail of experiments have increased, biologists have exhibited a growing interest and goal. Some biologists would like to follow the behavior of

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-33 Calculated sensitivities of nitrogen-vacancy centers to 1H nuclear magnetic moments quantified by the number of detectable 1H spins using a published formula.
SOURCE: Yu et al. 2021.

biomolecules, especially enzymes, in real time. The difficulty arises when one needs to increase the light intensity to levels required to probe the important mechanistic molecules, risking changing their behavior—or even damaging them. In some cases, ordered arrays of quantum sensors that are individually addressable and densely packed in a small volume are used to produce hyper-resolved images or maps that would far exceed the diffraction limit or even room-temperature super-resolution. Creating such an array requires a rigidly organized collection of qubits that are individually optically addressable. These arrays can be organized into a molecular structure, within a single crystal, or in a larger cross-braced scaffold of many solid-state quantum nanoparticles. In other NMR-related experiments, different isotopes such as deuterium, 13C, and 15N are used. Imaging spectroscopies sensitive to molecular vibrations can be used to detect deuterium but have low sensitivity for 13C or 15N. However, 13C and 15N have nuclear magnetic moments that can be detected in conventional NMR spectrometers, raising the intriguing possibility of using high-resolution and high-sensitivity quantum-enhanced magnetometry for detection. Such a tool would enable, for instance, maps of candidate drug distributions through different tissues and, perhaps, even which organelles the drug localizes to, or maps of protein synthesis upregulation in the brain in response to learning or memory formation. To achieve spatially dense measurements with high time resolution using quantum sensors, one will need to integrate the measurements with complementary classical sensing modalities that can provide nanoscale chemical information across an entire cell. These methods will exploit the exquisite sensitivity of enzymes to the local biochemical environment.

3.8.2 Quantum Light and Biological Processes

Other scientists interested in this difficult issue regarding the measurement or tracking of enzymatic activity in biological systems with low light intensities have invoked quantum metrology as a solution. In a proof of concept, conducted by Cimini and colleagues (2019), N00N states were employed to observe and measure enzyme

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-34 Examples of existing sensing protocols that leverage quantum mechanical properties such as coherence and entanglement to detect (a) electron–nuclear hyperfine interactions, (b) electric fields, and (c) space–time perturbations.
SOURCE: Yu et al. 2021.

activity. First, N00N states that are in circular-polarization basis are created from single photons with orthogonal polarizations (using SPDC) that underwent interference through a polarizing beam splitter (Figure 3-35). This interference is an important step because it generates photons in superposed states, hence exhibiting indistinguishable spatial and temporal degrees of freedom, which allows the creation of the N00N states. After the desired photon state is generated, it is then passed through a solution containing enzymes undergoing catalytic activity. In this example, the enzymes are hydrolyzing sucrose and producing mixtures of chiral products. Due to the chiral states, the molecule’s interactions with the photon creates a phase shift between the different N00N states. The extent of the phase shift is a function of the concentration of sucrose. Two avalanche photodiodes were used to measure these shifts. There are still challenges making this technique feasible for use (signal-to-noise detection ratio). However, this proof-of-principle work demonstrated researchers’ ability to study biocatalytic reactions to track enzymes’ activity in real time by taking advantage of different chiral systems and their distinct optical properties.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-35 (a) Schematic of experiment. (b) Two photons produced via spontaneous parametric down-conversion with orthogonal polarizations are sent on a polarizing beam splitter (PBS) to produce a N00N state. (c) 2φ between the two components of the N00N state is introduced. When reverting to the linear polarizations, the phase shift corresponds to a rotation by an angle φ/2 on both photons. (d) A half-wave plate (HWP) and a second PBS is used to analyze the polarization. The photons are recorded as coincidence counts based on detection using avalanche photodiodes.
SOURCE: Cimini et al. 2019.

3.9 USE BIO-INSPIRED QUANTUM PROCESSES TO DEVELOP NEW QUANTUM TECHNOLOGIES

Strategies to develop platforms for quantum sensing can be found within the realm of biology. Quantum biology is the study of the dynamics of quantum dynamical networks in the presence of an environment. It suggests that biological phenomena such as photosynthesis, enzyme catalysis, avian navigation, and olfaction may utilize fundamental quantum mechanical processes such as coherence, tunneling, and entanglement. Early suggestions for the role of quantum mechanics in biology were made by Schrödinger (2012) and Fröhlich (1968), who proposed that coherence may be the basis of biological oscillators. More recently, superconducting qubit architectures have been used to explore quantum mechanical models of photosynthetic light harvesting inspired by 2DES experiments (Potočnik et al. 2018). The debate over whether quantum physics applies to biological systems centers on the view that biological systems can be seen as noisy and complex from a quantum mechanical perspective (Tegmark 2000).

The suggestion that consciousness depends on biologically relevant coherent quantum processes and that these quantum processes correlate with neuronal synaptic activity (Hameroff and Penrose 2014) has been proposed. Calculation of neural decoherence rates finds that while decoherence timescales are ~10−13–10−20 sec, the relevant dynamical timescales (~10−3–10−1sec) for regular neuron firing are much longer, suggesting that the quantum coherence may not be relevant to neural network functions in higher-order organisms. Without a viable qubit specification, the role of quantum phenomena in higher-order neuronal processes remains controversial (Hameroff and Penrose 2014). Strong evidence for the role of coherence, entanglement, and superposition of states, however, has been found in biological photochemical reactions, sensory mechanisms (sight and smell), photosynthesis, and magnetoreception, thus providing fascinating models from biology as to how chemists might design and study quantum sensors for a range of environmental variables.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

The role of quantum phenomena in photochemical systems, particularly photosynthetic reaction centers, is perhaps the most well studied. Quantum coherence dictates energy transfer processes in the pigment–protein complex of chlorophyll with decoherence times of 600 fs (Engel et al. 2007; Ishizaki and Fleming 2009; Sarovar et al. 2010; Squire et al. 2013). Coherence in light-harvesting and electronic energy transfer processes in biological systems has been observed on femtosecond timescales, and the phenomena have been modeled to explain the process from classical physics to quantum superposition (Chenu and Scholes 2015).

Important sensory mechanisms such as those of sight and olfactory responses have been proposed to have a quantum origin. Vision depends on the sensing of photons in rods and cones, and sensitivities as high as one to three photons (single-photon counting) have been reported (Rieke and Baylor 1998), which is relevant as inspiration for the development of optical quantum sensors. The sensitivity of olfactory responses to vibrational or phononic processes of molecular systems, beyond simple structural factors, suggests that vibrational modes in molecular structures for olfaction may play a role in sensing through a mechanism analogous to inelastic electronic tunneling spectroscopy (Turin 1996).

Magnetoreception refers to the ability of organisms to utilize external magnetic fields to sense direction (Wiltschko and Wiltschko 2005, 2012). Magnetic field sensing can be with respect to the north or south lines of the geomagnetic field (magnetotaxis) (Rismani Yazdi et al. 2018) or with regard to field strength and inclination involving some combination of the geomagnetic and paleomagnetic field. Magnetoreception has been demonstrated in hundreds of organisms to date, from the orientation of swimming for magnetotactic bacteria, to the orientation of hives for bees and ants, to tunnel direction for mole rats, and to migration over long distances by many species of birds, sea turtles, and invertebrates such as mollusks and crustaceans. The mechanism of magnetoreception can be light dependent, as was found in many species of birds (the “avian compass”), or light independent (sea turtles, bacteria). The light-dependent mechanism is fundamentally a quantum phenomenon that relies on the quantum spin dynamics of transient photoinduced radical pairs that are formed via photoinduced electron transfer between a putative acceptor and donor in a relevant protein (Hiscock et al. 2016; Zoltowski et al. 2019). Coherent spin dynamics between singlet and triplet states are influenced by an external (geomagnetic) field, leading to changes in the quantum yield of the signaling state of the protein. Significant evidence suggests that the critical acceptor–donor pair is comprised of a tryptophan and flavin cofactor found in cryptochromes (Maeda et al. 2012) and the kinetics and quantum yields of photoinduced flavin–tryptophan radical pairs in cryptochrome were found to be magnetically sensitive in field strengths akin to Earth’s magnetic field (25–65 μT) (Maeda et al. 2008). The directionality associated with magnetic field detection, however, must include a contribution due to magnetic anisotropy in order to sense “direction” rather than just the magnitude of the Zeeman splitting of states. Neither the nature of the anisotropy (e.g., g-anisotropy and hyperfine, on a quantum mechanical level) nor how it is translated within a biological environment to “directionality” are understood. The light-independent mechanism of magnetoreception is a more classical mechanism; biomineralized magnetite Fe3O4 or greigite Fe3S2 sense changes in the direction and magnitude of external fields to initiate signal transduction pathways. The pathways are not well understood but are believed to involve mechanosensitive ion channels that translate the mechanical motion of the magnetic nanoparticles through the protein matrix to initiate signal transduction. Other magnetic sensing mechanisms based on electromagnetic induction have been evidenced in fish and sharks (Kirschvink, Walker, and Diebel 2001). The mechanisms provided through magnetoreception provide interesting models for the design of chemical systems for magnetic field sensing.

3.10 PROVIDE BROADLY ACCESSIBLE STATE-OF-THE-ART MEASUREMENT TECHNIQUES AND INSTRUMENTATION FOR THE CHEMISTRY COMMUNITY

3.10.1 Emerging Tools and Needs

As the field of chemistry grows into QIS, the tool set already developed for more traditional QIS experiments must be adopted and expanded. In particular, new tools would benefit the following categories: (1) single-molecule qubit measurement, (2) control and detection of ensembles of molecular qubits, and (3) advanced spectroscopy of entangled and classical interactions for molecular qubits. This section discusses these instrumentation classes as emerging tools and needs.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

3.10.1a Single-Molecule Qubit Measurement

Currently, molecular qubits and quantum systems are mostly studied as ensembles. While this approach is sufficient for measurements of dephasing or ensemble quantum states, as described in previous sections of this report, a need exists to measure quantum information for individual molecules, between individual parts of a molecule, and between a few molecule systems. The tools used for atomic and ion systems simultaneously allow the isolation of single particles or control over multiple particle systems and their interaction Hamiltonians. The ability to isolate quantum dynamics has led to significant scientific advances in atomic and ion systems, but a suitable technique has yet to be adapted for molecular qubits beyond a few atoms (Chou et al. 2017; Leibfried 2012; Lin et al. 2020; Loh et al. 2013; Shi et al. 2013; Wolf et al. 2016). A similar tool set would benefit molecular systems, where new interactions are possible by chemical synthetic control of qubits and control of multiple molecule systems.

It is important to develop methods for measuring single molecules within an ensemble. This category can be further broken down into the study and readout of the qubit or quantum sensors versus the characterization of molecular properties of the qubit or quantum sensor that lead to its successful operation. Molecular qubits that can be optically read out represent a relatively accessible measurement for most spectroscopy groups. Measuring single-molecule fluorescence is a well-established technique in chemistry and biological imaging (Lelek et al. 2021; Shashkova and Leake 2017). Following the development pathway of the NV center field, molecules can be dispersed on a surface and their fluorescence isolated using a variety of microscopy approaches (Scholten et al. 2021). This has been done, for example, with optically addressable molecular qubits, and a setup is shown in Figure 3-36 (Bayliss et al. 2020). However, further instrumentation development would help a synthetic chemist, for example, rapidly scan candidate molecules. The optics of Figure 3-36 would need to be simplified into a bench-top entangled spectrometer, which is an avenue worth exploring if rapid advances in chemistry for QIS are to be made.

Isolating and measuring microwave active molecular qubits is a more challenging task because the microwave wavelength (millimeter and longer) is far from that of visible light (nanometer) and is incommensurate with the size of molecules. A need exists to adapt current atomic resolution methods to be suitable for molecular qubits. Atomic resolution instruments fall into two main categories: scanning probe microscopy and electron imaging techniques. While electron imaging methods like transmission electron microscopy (TEM) have the atomic resolution needed to characterize even small molecules (through cryo–electron microscopy and microcrystal electron diffraction) (Jones et al. 2018), it is more challenging to extract quantum-state information; however, future developments like Lorentz TEM of magnetic fields could prove transformative. On the other hand, scanning probe microscopies, which include instruments like an atomic force microscope and an STM, are already used to measure single-molecule magnetic properties as well as to read out their spin states (see Choi [2019] and Moreno-Pineda and Wernsdorfer [2021] for a review).

Scanning probe techniques can also be adapted to perform nanoscale optical and microwave experiments. Generically speaking, near-field scanning probe microscopes use resonant apertures of various forms, such

Image
FIGURE 3-36 Experimental setup for characterizing all optical qubits.
SOURCE: Bayliss et al. 2020.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-37 Different types of scanning near-field optical microscopes: (a) aperture scanning near-field optical microscopy with angular resolved detection, (b) apertureless configuration, and (c) scanning tunneling optical microscope.
SOURCE: Hecht et al. 2000.

as metal tips, to localize an incident electromagnetic field to a sub-wavelength regime (Figure 3-37) (Anlage et al. 2001; Farina and Hwang 2020; Hecht et al. 2000; Mauser and Hartschuh 2014; Rosner and van der Weide 2002; Seo and Kim 2022). Common abbreviations for the techniques are near-field scanning optical microscopy for visible light to terahertz characterization and near-field scanning microwave microscopy for the long-wavelength microwave version. The main difference is the source: visible to terahertz light can be generated by nonlinear frequency generation from a laser, whereas microwaves are created using electronic signal generators and antennas or waveguides when needed. Scanning near-field techniques can reach tens of nanometers in resolution to isolate single molecules or look at interactions between nearest-neighbor molecules. The methods are also easily converted into transient measurements, wherein an optical or microwave pump is used to initiate an excited state of the molecular qubit and then another wavelength region is used as the probe. This allows for measurements of the underlying dephasing mechanisms that are controlling the molecular magnet or qubit operation.

While not as elegant as ion trap measurements, scanning near-field microscopies will be an emergent tool for molecular QIS systems. Multiple companies now sell instruments adaptable for chemistry in QIS work. Depending on the complexity and options selected, the instruments range from the hundreds of thousands to the million-dollar range. The instruments also have multiple applications and therefore can be invested as user facilities. Cutting-edge methods that use STM feedback circuits are now also allowing for sub-nanometer resolution, even with microwave frequencies, and are emergent tools that allow true quantum readout between different spin centers of a molecule (Imtiaz, Wallis, and Kabos 2014). Furthermore, by combining scanning probe microscopy methods with entangled photon spectroscopy, how molecules control or rely on entanglement can be directly measured on the atomic scale.

Combined, the application of single-molecule characterization tools to chemistry in QIS can (1) measure the quantum state of a molecular qubit after being initialized and its decoherence time; (2) use steady-state and ultrafast characterization techniques to measure spin-lattice, spin-spin, and other interactions to understand decoherence mechanisms; and (3) repeat these measurements with nearby molecular qubits to understand coupling mechanisms. These types of measurements would give clues as to how to delineate and achieve multiple qubit interactions in molecular quantum sensors and other QIS systems.

3.10.1b Control and Detection of Ensembles of Molecular Qubits

The application of molecular qubits in quantum scenarios requires the delineation of interactions in an ensemble. Multiple molecular qubits, molecular magnets, or molecular sensors will need to be addressed individually and measured in a complicated environment. While some differentiation can be achieved through synthetic means, as demonstrated by sensing with diamond and other vacancy centers, emerging instrumentation will also play a dominant role.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-38 Schematic of making individual spin qubits controllably interact through the use of an optical lattice approach.
SOURCE: NIST 2009.

First, the topic of constructing defined ensembles of molecular qubits will be discussed. Synthetic methods discussed in Chapter 2 can be used to specifically anchor multiple qubits in ordered fashions; however, more artificial instrumental methods like optical trapping and scanning probe microscopies should also be mentioned. Scanning probe microscopy, the base technique from the previous section, has famously been used to place and order atoms and molecules on a substrate (Gross et al. 2005; Hla and Rieder 2003), even to the point where molecular logic gates have been built (Leisegang et al. 2021). While a tedious technique, copious research has been put into this topic, and artificially creating arrays of molecular qubits could provide a fruitful equivalent to an atomic ion trap. Optical forces can also be used for molecular manipulation. For example, optical tweezers take advantage of the response of a dielectric sphere illuminated by laser light for high levels of precision of molecular control (Bustamante et al. 2021; Moffitt et al. 2008). Applications of optical tweezers were discussed in Section 2.2.2. The downside of this approach is that the molecules must be attached to dielectric spheres in order to provide levitation or multiple ordering. Optical lattices aim to overcome this issue by using complex arrays of interfering laser beams. They are commonly used in atomic, ion, and few-atom molecule work in QIS, although larger molecules are still a technological frontier (Chou et al. 2017; Leibfried 2012; Lin et al. 2020; Loh et al. 2013; Shi et al. 2013; Wolf et al. 2016). Instead of using a dielectric sphere, the system is cooled to ultralow temperatures using a laser or similar methods and then patterned using a complex array of beams with constructive and destructive interference (Figure 3-38). Both optical tweezers and lattices can be considered critical to chemistry in QIS growth because they allow a more free-form and adjustable ordering than is possible with surface manipulation.

Second, instead of trying to create arrays of single molecules, an alternative approach is to isolate a specific qubit(s) within an ensemble. This is especially important for quantum sensing, one of the recommended research priorities of the report, where multiple near-identical qubits will be present in complex physical and biological environments. Inspiration can again be taken from the field of biological imaging, which already addresses this exact challenge. Complex biological microscopes are often already available in commercial forms at campus user centers and are therefore interesting to explore. For example, multiphoton imaging uses nonlinear optical properties to create three-dimensional (3D) images with molecular tags and tagless technologies. Multiphoton microscopy works by overlapping two or more pulsed laser beams within a sample (Hoover and Squier 2013; Lecoq, Orlova, and Grewe 2019). When overlapped in time and space, a multiphoton process, such as two-photon fluorescence, can occur. As the beams are moved throughout a sample, a 3D image is produced, often with up to 1 mm penetration depths and hundreds of nanometers’ resolution (Figure 3-39). Light sheets can also be used for rapid, real-time image acquisition. When combined with super-resolution and stimulated depletion techniques, tens of nanometers’ spatial resolution is possible (Moneron and Hell 2009). An emergent area of research would be adapting these techniques to isolate and optically address molecular qubits. This would require instrumentation development as well as the creation of appropriate molecular qubit tags that respond to nonlinear or multiphoton activation.

Alternatively, instead of designing qubits to react to multiphoton excitations, tagless methods like stimulated Raman scattering are also now reaching the <100 nm spatial resolution barrier (Li et al. 2021; Qian et al. 2021;

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-39 A rapid optical clearing protocol using 2,2′-thiodiethanol for microscopic observation of a fixed mouse brain. Multiphoton microscopy techniques developed in biology could potentially be used for addressing spin qubits in an array or for measuring multiple molecular qubit sensors in an environment.
SOURCE: Aoyagi et al. 2015.

Shi, Fung, and Zhou 2021). In this case, a molecular qubit would be imaged by its unique Raman spectral profile. Since Raman scattering can measure individual bonds, methods may be developed that can measure decoherence indirectly by vibrational dynamics or, more challengingly, directly through their fine structure. Other nonlinear microscopy techniques, like second and third harmonic generation or sum frequency generation, are also sensitive to local polarizations and vibrational populations and can be used as label-free methods for molecular qubit characterization and interaction (Lim 2019; Rehberg et al. 2011; Wang and Xiong 2021).

3.10.1c Advanced Spectroscopy of Entangled and Classical Interactions for Molecular Qubits

Previous sections have focused on ways to measure or control individual molecular qubits or networks thereof. However, instrumentation development on the ensemble level presents an opportunity to help understand how molecular qubits operate and transduce quantum information. Although microwaves have been used for the control and readout of most molecular qubits to date, how a molecule decoheres and its quantum coupling are often explored using ultrafast lasers. As mentioned throughout this report, techniques like sum frequency generation and stimulated Raman scattering can be used to understand time domain information about bond vibrations; the optical Kerr effect can be used to measure spin polarizations; and multidimensional spectroscopy (2DIR and 2DES) or transient EPR can be used to create and measure molecular spin center coupling within a molecule.

Further development of existing nonlinear optical techniques to operate in conjunction with magnetic fields, microwave access, and optical readout of spin will advance the field. Conventional pulsed EPR cannot measure rapid electron spin relaxation (less than ~5 ns) and does not provide the spatial resolution of optical techniques. Ultrafast spin relaxation processes have been studied in spin-½ systems using all-optical methods such as Faraday rotation to measure optical induction of ground-state magnetization dynamics and subsequent observation of quantum-beat free-induction decay (Furue et al. 2005). While long T1 and T2 times are required for quantum information processing applications, measurement of ultrafast spin dynamics still enables fundamental insight into molecular spin–phonon coupling and decoherence mechanisms. A particularly powerful way to explore these mechanisms that need to be developed is probing variable-temperature, variable-field regimes that are typically inaccessible to EPR and AC magnetometry.

Although ensemble based, these techniques remain prominent because high-purity synthesis products can be created. The methods of this section are more mature and accessible, but continued development could further simplify the tools to the level that a synthetic laboratory could use them for rapid characterization without needing a laser spectroscopist collaborator. While these collaborations are highly encouraged and productive, they create a

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

barrier for universities and laboratories that do not have the funds for complex laser tools. This report concludes that the development of new specialized techniques is needed, but simultaneous efforts to simplify and lower the cost of existing nonlinear or entangled spectroscopy techniques are critical for accelerating molecular qubit design rules. These developments will also increase ready access to user facilities housing the needed tools.

3.11 IMPROVE INFRASTRUCTURE FOR CHEMICAL MEASUREMENTS IN QIS

Physical resources are required to support the range of established and emerging tools for the experimental characterization of quantum behavior discussed in this chapter. This section reviews available infrastructure in the United States today and makes recommendations for the prioritization of future resources to advance the field of QIS.

This section divides user facilities into three main categories that are based on the scale of the infrastructure. Box 3-1 briefly defines these categories as they are used in the remainder of this section. Although the National Science Foundation (NSF) has established funding ranges (in dollars) that are defined using similar terms, the committee acknowledged the continuously changing costs of various techniques and elected not to assign dollar values to these scale definitions.

Another key element of establishing and maintaining a strong infrastructure for QIS chemistry research in the United States is workforce development. The cultivation of an educated, well-equipped human resources pipeline to support the maintenance and operation of user facilities is discussed in Chapter 5.

3.11.1 Laboratory-Scale, Single–Principal Investigator Centers

A need exists to develop chemistry-oriented QIS laboratories if the field is going to progress. As of now, most single–principal investigator (PI) laboratories do not have access to the basic tools needed to characterize microwave qubit interactions, such as transient EPR, nor the ability to characterize optical qubits. Most chemistry in QIS research has been accomplished by the partnering of synthetic chemistry laboratories with physics or physical chemistry laboratories that specialize in quantum optics or transient EPR techniques. However, these collaborations do not exist at every university and present a barrier to development. Standardized and commercial instruments need to be developed for basic qubit characterization, especially for measuring properties like coherence times or spin-coupling interactions in multiqubit architectures, to allow any synthetic laboratory access to chemistry in QIS research.

Even for quantum optics and optics laboratories, a need for investment remains. Magnetic field and microwave generators are not commonly associated with ultrafast laser optics, and, at the same time, advanced optical capabilities are rarely integrated with transient EPR spectrometers. These capabilities, along with cryostats, represent significant investments. Physical chemists have developed a broad range of nonlinear and linear spectroscopies that can measure almost all of the details of the molecular interaction Hamiltonian. However, these techniques are not yet optimized for measurements of molecular qubits. Ultracold atoms and optical lattice research is currently focused in the physics community, and extension to molecular systems beyond a few atoms will enhance this area of research greatly.

While custom instrument development is needed, multiple tools can already be purchased commercially and adapted to the chemistry in QIS tasks. For example, the multiphoton and tagless Raman imaging microscopes

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

developed for biology can be used to isolate and measure molecular quantum sensors in complex environments. Optical tweezer systems can be purchased for isolating single-molecular qubits and exploring factors like strain and stiffness on the molecular qubit’s dephasing time. Near-field scanning probe microscopies like NSOM and NSMM have reached maturity and can be used for single- and multiple-qubit interaction studies with atomic resolution. Scanning probe methods can also be used to arrange molecules on surfaces. Transient EPR is a dominant tool that is commercially available, albeit at a high cost and on a limited basis. All of these instruments represent significant investments in the hundreds of thousands to several-million-dollar range, but they can be leveraged for multiple experiments across multiple departments to help their adoption.

Driven by communications industries, rapid advances in microwave technologies are such that the hardware employed in transient EPR spectrometers is becoming more advanced, more widely available, and more affordable. Indeed, perhaps with the exception of the magnet, the components needed to assemble quite advanced transient EPR spectrometers can be acquired at a fraction of the current cost of a fully assembled turnkey instrument. Of course, most research groups do not possess the technical expertise to build their own transient EPR spectrometers. Nevertheless, the increased need for these capabilities motivates creative solutions to the significant financial barrier that currently limits the availability of such systems.

There are encouraging signs that several new startups, often supported by Small Business Innovation Research–type federal funding, are entering this scientific space. In turn, this is leading to exciting new routes to reduce costs such as Open Source Hardware,1 where all relevant information required to produce a given hardware item can be posted and made available to the wider research community at no cost. Indeed, one can already find examples of extremely low-cost components relevant to transient EPR spectroscopy. Another area of rapid progress is the development of advanced on-chip EPR spectrometers that employ tiny permanent magnets (Hassan et al. 2021, 2022; Künstner et al. 2021; Lotfi et al. 2022), with the main spectrometer hardware often having a footprint no larger than a thumb drive. Such systems offer the added advantage of exceptional sample sensitivity. Although still developmental, these instruments are starting to demonstrate some of the capabilities of commercial spectrometers and could soon become commercially available at low cost. Therefore, the prospects for more widespread access to transient EPR capabilities over the next 10 years are promising and should be encouraged by funding agencies, although it is clear that these capabilities will remain somewhat limited in comparison to the most advanced systems (see below). It will be important to ensure that issues related to precision, reproducibility, and safety are addressed if such low-cost instruments become more widely available. However, it should be noted that the international EPR community successfully polices such issues upon the emergence of any new hardware or methodology.

3.11.2 Mid-Scale University User Facilities and Centers

While the prospects for low-cost hardware described in the previous section are exciting, there will always be a need for high-end instrumentation, as well as one-of-a-kind spectrometers offering truly unique capabilities, developed by the leading instrument builders. State-of-the-art commercial spectrometers are essential for benchmarking promising molecular systems and enabling measurements (e.g., of coherence and weak spin–spin interactions) that are simply not possible by other means (Kundu et al. 2023). As noted in a recent article in the International EPR (ESR) Society newsletter (Gallez 2022), “Those universities with modern pulsed EPR systems are currently swamped with demand.” However, there are significant barriers to accessing instrumentation in the leading EPR laboratories. Typically, these instruments are operated in support of targeted research programs with relatively narrow scope, and access is limited to collaborators on these projects. The best way to remove these barriers is to locate high-end instrumentation and promote the development of new methodologies (e.g., those discussed in Section 3.11.3) at national user facilities. These organizations are truly open to the wider research community through merit-based proposal processes that are reviewed externally. They support world-leading staff scientists who can optimize the measurement time even for the most inexperienced user (from running the measurements to the data analysis, interpretation, and drafting of manuscripts); they educate their users, thus greatly

___________________

1 The website for Open Source Hardware is www.oshwa.org, accessed April 27, 2023.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

expanding intellectual resources within the wider chemistry community; and they enforce policies related to, for example, FAIR data principles.

Although several EPR facilities are scattered across the United States (see Figure 3-40), most of them are supported by the National Institutes of Health, and their mission is focused on biomedical, biochemical, and biophysical research (e.g., Cornell [ACERT]; University of Wisconsin, Milwaukee [National Biomedical EPR Center]; Miami University in Ohio [Ohio Advanced EPR Laboratory]; and the University of Denver [EPR Center]). These facilities generally are not open to chemists working on QIS problems, even though a handful of experiments have been performed using instruments located at these centers (Zadrozny et al. 2017). The Department of Energy also supports facilities focused on environmental and energy-related research (e.g., Argonne National Laboratory, National Renewable Energy Laboratory, Pacific Northwest National Laboratory, and University of California, Davis). The only U.S. EPR user facility that currently supports research at the interface between chemistry and QIS is the NSF Division of Materials Research–funded NHMFL, where the majority of experiments are conducted at magnetic fields and frequencies that are considerably higher than those of commercial instruments (Baker et al. 2015; Morley, Brunel, and van Tol 2008). However, users of the facility also have access to commercial X- (9–10 GHz) and W-band (94 GHz) transient spectrometers as well as a high-power W-band transient spectrometer developed at St. Andrews University (Cruickshank et al. 2009), resources that are highly valued by the user community. The facility is also staffed with internationally renowned EPR support scientists. Therefore, it serves as a model facility for carrying out research on molecular spin qubits of the kinds discussed in Chapter 2, with more than two-thirds of its users drawn from the chemistry community. It is a resource not only for U.S. scientists but also for international researchers, with about 50 percent of its users coming from overseas—particularly from Europe, where a large and active chemistry QIS community already exists. Indeed, the strong demand from overseas users is a testament to the world-leading stature of the NHMFL EPR facility, as similar user centers exist in Europe and Asia. The downside is that limited resources (both in terms of staffing and instrumentation) are making it more difficult for the facility to keep up with user demand, with considerable growth in activity deriving from chemistry-inspired QIS research.

Image
FIGURE 3-40 Locations of electron paramagnetic resonance (EPR) groups, centers, and user facilities in the United States. The pins mark National Institutes of Health (NIH) and Department of Energy (DOE)–funded user facilities that primarily support biochemical and biomedical research. The only major user facility directly supporting QIS research is the National Science Foundation (NSF)–funded National High Magnetic Field Laboratory in Florida. The solid circles denote locations with active EPR research groups.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

It is vital for U.S. competitiveness in the growing research area at the interface between chemistry and QIS that the NHMFL and related user facilities in the United States remain at the cutting edge and not lose ground to similar centers overseas. To do so will require continual renewal of instrumentation at these centers. This applies not only to the major infrastructure (e.g., magnets and beamlines; see Section 3.11.3) but perhaps more importantly to the user end-stations (i.e., mid-scale instruments), where rapid technological advances can lead to the situation where a newly developed capability can become uncompetitive (or even obsolete) within a time frame of just a few years. A perfect example is microwave amplifiers and sources where EPR experiments performed today simply were not possible just 10 years ago (Subramanya et al. 2022). Maintaining adequate staffing at these facilities is also vital because of the critical role these people play in much of the research output and the education of users, including junior and early career scientists. Indeed, staff expertise is one of the key factors that fuels the world-leading status for many of the major user facilities in the United States, even though staff are often stretched very thin. Meanwhile, increased efforts should encourage the translation of novel capabilities developed in single-PI laboratories to user facilities, where they can be made available to the wider chemistry community via the knowledgeable staff at these centers. Examples include optical excitation and detection, and integrated scanning probe techniques.

Aside from major user facilities, a strong case can be made for establishing regional user hubs that support mid-scale instrumentation, which the chemistry QIS community relies on for advancing its research—for example, instrumentation that would be out of reach for all but the most research-intensive universities. Indeed, the natural place to locate these hubs would be major R1 academic institutions that have a critical mass of faculty and researchers with relevant expertise. However, it would also be important to staff these hubs with independent research scientists who would primarily serve the users of the instrumentation, in much the same way that a university funds staff to support campus instrumental needs. The hubs would need regular oversight by external entities in order to avoid situations in which the institution absorbs much of the added instrumental capability in support of its own interests. One could imagine national umbrella organizations that handle user applications to the national facilities and regional hubs, a model already employed among the European network of magnet laboratories.2 National facilities already know how to run user programs, and their insight and oversight could be a model for success of a distributed network of user hubs. This model would alleviate some of the increasing demand at existing national centers, while allowing these major facilities to focus on their core mission (e.g., a focus on high-field measurements at the NHMFL). The addition of hubs would also allow for anticipated growth in research at the interface between chemistry and QIS.

Finally, it is vital to U.S. competitiveness in the area of chemistry QIS research that a steady funding stream exists for the development of truly unique methodologies that overcome the limitations of commercial instrumentation. Given the explosive growth in research at the QIS and chemistry interface, the need for such new developments is becoming increasingly urgent, to the extent that it represents a rate-determining step toward future progress. As an example, most of the current work on spin qubits focuses on single-molecule coherence, whereas the next frontier clearly involves the exploration of entanglement and potential implementation of quantum gates involving multiple qubits (Chiesa et al. 2020; Ferrando-Soria, Pineda et al. 2016; Godfrin, Thiele et al. 2017; Hussain et al. 2018; Ullah et al. 2022). Short-term needs (on a timescale of three to five years) for realizing much of the vision laid out in this chapter are clear. Although recent advances in high-end commercial spectrometers have been breathtaking, such instruments are designed with a broad range of applications in mind (i.e., they are not optimized for specialized chemistry QIS tasks). In particular, bandwidth limitations restrict EPR studies to relatively narrow frequency ranges centered on just one of the various microwave bands (e.g., the L-, S-, X-, and Q-bands at 1, 3, 9, and 34 GHz, respectively). Because of the rapid development of microwave technologies, wideband NMR-type EPR experiments are now becoming routine at frequencies in the 10–50 GHz range, with experiments up to 400 GHz not lagging far behind (Subramanya et al. 2022). As noted in Section 3.1.3, instruments are urgently needed that combine transient EPR and NMR capabilities at multiple widely separated frequencies (from tens of megahertz to hundreds of gigahertz) with optical and other modes of excitation and detection. This will enable qubit initialization and excitation of multiple electron and nuclear spin transitions either within the

___________________

2 The website for the European Magnetic Field Laboratory is https://emfl.eu, accessed April 28, 2023.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

same molecule (e.g., a qudit) or among multiple coupled spin centers, thereby facilitating execution of multiqubit gate operations, initially among molecular ensembles. In the longer term (5 to 10 years), efforts would be directed toward arrays of molecules on surfaces. The costs for such developments will be considerable, likely requiring new source/amplifier technologies that extend beyond the current 5G frequency band. Therefore, design criteria must be carefully formulated and based on a viable chemical approach, involving teams of instrument builders, engineers, and synthetic QIS chemists. Significant benchmarking of proposed molecular species represents an important first step in the design process so that future instrument developments do not fail simply because the molecules do not perform as expected. The scale and complexity of these development projects are comparable to those typically associated with national facilities. Moreover, staffing requirements, both during development and upon deployment, suggest that these projects be managed through national facilities. This will ensure availability of necessary expertise, project management, and maximum eventual access.

3.11.3 Large-Scale User Facilities

The committee did not identify any need for new large-scale national facilities. However, barriers to access were clearly identified. Current facilities are oversubscribed, and clearly a need exists to support efforts that increase access—for example, by creating regional instrumentation hubs that relieve national facilities of some of their more routine user demand, increasing availability of magnet/beam time at national facilities to the chemistry QIS community, and developing new beamlines and resonance magnets. However, as noted in the previous section, available user end-station instrumentation also represents a major current barrier to access at national facilities. Therefore, a steady funding stream is needed to develop and sustain support for chemistry QIS instrumentation at national facilities, as has been the case for other large-scale entities such as synchrotrons.

3.11.3a Magnet Laboratories

As noted in earlier sections of this report, strong spin–orbit coupling and ligand field effects can give rise to situations where metal-based spin qubits are EPR-silent at the conventional low frequencies (<50 GHz) and fields (<1.5 T) employed in commercial spectrometers, thus necessitating measurements at high magnetic fields and at frequencies stretching into the terahertz and far-IR regimes. The same is true for NMR, where the most advanced spectrometers capable of targeting quadrupolar nuclei spanning much of the periodic table operate at magnetic fields in excess of 30 T (Gan et al. 2017). For this reason, a large body of work in the molecular QIS field has been performed at large-scale user facilities such as NHMFL.

Much of the high-field research performed today involves powered (resistive) magnets, and the power supplies ultimately limit the amount of time available to the user community at the very highest fields. The current situation at NHMFL is that costs prohibit 24/7 operations. Therefore, some scope exists for increasing user operations without the need for significant investment in infrastructure, although the associated staffing and electricity costs would be quite considerable. Recently, however, the laboratory successfully brought online the first all-superconducting 32 T user magnet, with high-temperature superconducting inner coils (Weijers et al. 2016). Compared to resistive magnet technologies, it provides a lower noise environment for precision measurements, as well as a larger sample space. This magnet is now part of the user NHMFL program, although it is not currently instrumented for EPR or solid-state NMR experiments. Even though this magnet runs independently of the high-power resistive magnets (that require a 56 MW supply), it would not be easy to operate outside of the facility due to a very high helium consumption and complex supporting infrastructure (not to mention that this technology remains developmental). The resonance users of the facility (both EPR and NMR) recognize the tremendous opportunities that such a magnet provides in terms of increased access to the very highest magnetic fields. The current version does not provide the homogeneity needed for high-resolution solid-state NMR, but it is more than adequate for most EPR applications as well as certain specialized NMR applications (and further development of methodologies such as DNP). The development of a second all-superconducting magnet dedicated to resonance applications as well as the future development of high-resolution all-superconducting NMR magnets would serve as a very significant research resource for the chemistry QIS community.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Of course, the availability of new magnets and increased magnet time is not sufficient by itself. Investment in EPR and NMR hardware optimized to these high-field magnets would increase research in chemistry and QIS. Such resources historically have been severely limited within NHMFL’s core funding, with most recent new developments being funded through separate grant applications. These activities impose a significant workload on staff, which either subtracts from user support time or imposes barriers that prevent staff from pursuing such development opportunities. Another area of opportunity is increasing involvement of the user community to fund such developments. However, the nature of the user community is often such that users lack the expertise to lead such an effort, instead contributing only to the scientific case. But this shifts the workload back to the scientific staff at the laboratory. This and the obvious limitations in the number of awards made to a given institution currently represent a significant barrier to the development of new experimental hardware for magnetic resonance. Most of this hardware falls under the mid-sale category and is discussed in the previous section. However, we discuss this issue again here, as it clearly represents a major barrier to progress in research at the interface between chemistry and QIS, and also limits access to large-scale facilities.

3.11.3b Light Sources and Synchrotrons

Free electron lasers and synchrotrons are regularly used for the characterization of molecular structure and dynamics (Fukuzawa and Ueda 2020; Heinz et al. 2017; Young et al. 2018). The same approaches apply to molecular qubits, particularly since transient and steady-state X-ray measurements are element-specific and can separate, for example, a metallic spin center from its ligands. Spin sensitivity is also possible through magnetic dichroism and related experiments. However, the development of novel beamlines for molecular qubit research is needed as well.

The first category revolves around the modification of transient absorption and resonant inelastic X-ray scattering beamlines so that they are capable of low-temperature and microwave experiments. Most transient measurements use optical to terahertz pulses, which can excite electronic and vibrational states but not initiate molecular qubits. Using a transient microwave pulse to initialize the qubit and an X-ray pulse to measure the spin-dependent polarization and coupled lattice dynamics would be particularly powerful. Such a beam station could handle molecular jets, liquids, or thin-film samples. Near-field X-ray microscopy using zone plates has also progressed significantly, and its element-specific images could be used for investigating coupled or individual molecular qubits (Shapiro et al. 2020).

The second category of instrumentation revolves around the creation of beamlines to explore how entangled X-rays can be used for novel spectroscopy or characterization of qubits (Defienne et al. 2021; Durbin 2022; Röhlsberger, Evers, and Shwartz 2014; Smith, Wang, and Shih 2020). Similar to the previous section on nonlinear entangled photon spectroscopy (Section 3.4.1), researchers predict that X-rays can linearize multiphoton processes, increase SNRs, and be used for novel “ghost” imaging protocols. This would allow lower-power experiments, solving one of the biggest issues with nonlinear X-ray experiments: sample damage.

For entangled X-ray experiments to be feasible, more efficient sources and optics need to be developed (Röhlsberger, Evers, and Shwartz 2014; Volkovich and Shwartz 2020). Entangled photons are created by SPDC, in which a pump photon couples to a vacuum state to be split into two entangled photons whose energy and momentum sum to that of the pump photon. X-ray processes have low nonlinear coefficients because the refractive index and nonlinear coefficients are near unity. While various nonlinear processes like second harmonic generation, sum frequency generation, and TPA have been achieved using X-ray free-electron lasers, entangled photon generation through SPDC has a million to billions of times lower conversion efficiency than these nonlinear processes. In the hard X-ray regime, phase matching occurs through the nonlocal electron plasma and lattice periodicity even in centrosymmetric materials. Current entangled X-ray photon generation methods rely almost exclusively on diamond for phase matching with conversion efficiencies in the 10−12 range (Figure 3-41) (Shwartz et al. 2012).

However, recent reports of using periodically poled materials (similar to visible light entangled photon generation) have shown promising efficiencies in the 10−6 range (Shapiro et al. 2020). Metal and dielectric stacks are already employed in extreme ultraviolet optics, so the technology is already in place. For example, using polar materials, extreme ultraviolet (<100 eV) SPDC was recently measured to be one million times more efficient than hard X-rays (Sofer et al. 2019). Free-electron lasers can also natively generate entangled photon and

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Image
FIGURE 3-41 The generation of X-ray entangled photons by spontaneous parametric down-conversion in diamond has been demonstrated at keV energies at 10−12 efficiencies.
SOURCE: Shwartz et al. 2012.

electron–photon pairs (Wong and Kaminer 2021). Given that specialized detectors and long experimental times are needed for most X-ray experiments, the creation of dedicated beamlines could revolutionize nonlinear X-ray source creation as well as exploration of the use of entanglement.

3.11.4 Limited Natural Resources

Applications of rare-earth and actinide elements in the area of optically addressable qubits are discussed in Section 2.2.2 of this report. An important consideration in many of these examples is the availability of naturally occurring isotopes with a variety of nuclear spin quantum numbers, I. A perfect example is the I = ½ 171Yb+ ion that is employed in ion trap quantum computers (Wright et al. 2019), where the hyperfine interaction involving the lone unpaired 6s electron gives rise to a hydrogenic-like situation and a massive 12.6 GHz clock transition. Similar targets may be anticipated in pursuit of ideal molecular qubit candidates, necessitating isotopic purifications or the production of particular isotopes at nuclear reactor facilities. Access to such isotopes will likely be limited and costly, with sub-milligram syntheses precluding many standard characterization techniques that lack sensitivity. Sample recycling will therefore be essential. Such approaches are already employed—for example, in QIS studies of endohedral metallofullerenes, which are typically extracted and purified via chromatographic techniques. Indeed, several gadolinium endohedral metallofullerenes have recently shown great promise as potential molecular qudits (d-dimensional qubits) (Fu et al. 2022; Hu et al. 2018), where the fullerene cage provides added protection from environmental decoherence sources.

Finally, the ongoing helium crisis represents a very significant challenge for researchers working in the QIS field, where many studies are conducted at temperatures in the sub-10 K range. Moreover, many groups continue to rely on older liquid helium–based cryogenic systems. It goes without saying that continued efforts must be pursued to reduce reliance on such cryogenic approaches, by furnishing researchers with closed-cycle refrigeration systems.

3.12 SUMMARY OF RESEARCH PRIORITIES AND RECOMMENDATION

The following fundamental research priorities have been identified by the committee and extensively discussed in Chapter 3 as those that the Department of Energy and NSF should prioritize within the target research area of “measurement and control of molecular quantum systems.”

Research Priorities:

  • Develop new approaches and techniques for addressing and controlling multiple electron and nuclear spins and optical cycling centers in molecular systems.
  • Develop techniques to probe molecular qubits at complex interfaces to inform their systematic control.
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
  • Develop enhanced spectroscopic and microscopic techniques by creating
    1. entangled photon sources with higher yield and better spectral coverage, and
    2. high-finesse cavities and nanophotonics for molecular qubit systems.
  • Develop and exploit alternative approaches to spin polarization and coherence control (e.g., chirality-induced spin selectivity and electric field effects).
  • Use molecular systems to teleport quantum information over distances greater than 1 μm with high fidelity.
  • Develop molecular quantum transduction schemes that take advantage of entangled photons as well as entangled electrons and nuclear spins.
  • Advance quantum sensing techniques to further understand biological systems.
  • Use bio-inspired quantum processes for the development of new quantum technologies.

In sum, in this chapter the committee discussed infrastructure and instrumentation needs at various levels. The committee provides Recommendation 3-1 to ensure that the infrastructure needs and instrumentation support required to drive chemistry QIS research in the United States are met on a competitive timescale.

Recommendation 3-1. The Department of Energy and the National Science Foundation should support the development of new instrumentation and techniques for the unique needs at the interface of chemistry and quantum information science. Broader access to laboratory-scale and mid-scale instrumentation is needed for the field to progress. For example, investments should be made in time-resolved magnetic resonance and optical spectroscopy. Support is required for professional staff to train users in the operation and utilization of these instruments, as well as to address new technique development and maintenance needs.

REFERENCES

Aguilà, D., L. A. Barrios, V. Velasco, O. Roubeau, A. Repollés, P. J. Alonso, J. Sesé, S. J. Teat, F. Luis, and G. Aromí. 2014. “Heterodimetallic [LnLn′] Lanthanide Complexes: Toward a Chemical Design of Two-Qubit Molecular Spin Quantum Gates.” Journal of the American Chemical Society 136(40):14215–14222. doi.org/10.1021/ja507809w.

Ahn, W., F. Herrera, and B. Simpkins. 2022. “Modification of Urethane Addition Reaction via Vibrational Strong Coupling.” ChemRxiv. Cambridge: Cambridge Open Engage. doi.org/10.26434/chemrxiv-2022-wb6vs.

Allodi, G., A. Banderini, R. De Renzi, and C. Vignali. 2005. “HyReSpect: A Broadband Fast-Averaging Spectrometer for Nuclear Magnetic Resonance of Magnetic Materials.” Review of Scientific Instruments 76(8):083911. doi.org/10.1063/1.2009868.

Altenhof, A. R., A. W. Lindquist, L. D. D. Foster, S. T. Holmes, and R. W. Schurko. 2019. “On the Use of Frequency-Swept Pulses and Pulses Designed with Optimal Control Theory for the Acquisition of Ultra-Wideline NMR Spectra.” Journal of Magnetic Resonance 309:106612. doi.org/10.1016/j.jmr.2019.106612.

Anderegg, L, B. L. Augenbraun, Y. Bao, S. Burchesky, L. W. Cheuk, W. Ketterle, and J. M. Doyle. 2018. “Laser Cooling of Optically Trapped Molecules.” Nature Physics 14(9):890–893. https://doi.org/10.1038/s41567-018-0191-z.

Anderegg, L., B. L. Augenbraun, E. Chae, B. Hemmerling, N. R. Hutzler, A. Ravi, A. Collopy, J. Ye, W. Ketterle, and J. M. Doyle. 2017. “Radio Frequency Magneto-Optical Trapping of CaF with High Density.” Physical Review Letters 119(10):103201. doi.org/10.1103/PhysRevLett.119.103201.

Anderegg, L., L. W. Cheuk, Y. Bao, S. Burchesky, W. Ketterle, K.-K. Ni, and J. M. Doyle. 2019. “An Optical Tweezer Array of Ultracold Molecules.” Science 365(6458):1156–1158. doi:10.1126/science.aax1265.

Anlage, S. M., D. E. Steinhauer, B. J. Feenstra, C. P. Vlahacos, and F. C. Wellstood. 2001. “Near-Field Microwave Microscopy of Materials Properties.” In Microwave Superconductivity, edited by H. Weinstock and M. Nisenoff, NATO Science Series, Vol. 375. Dordrecht: Springer. https://doi.org/10.1007/978-94-010-0450-3_10.

Aoyagi, Y., R. Kawakami, H. Osanai, T. Hibi, and T. Nemoto. 2015. “A Rapid Optical Clearing Protocol Using 2,2′-Thiodiethanol for Microscopic Observation of Fixed Mouse Brain.” PLoS ONE 10(1):e0116280. doi.org/10.1371/journal.pone.0116280.

Ardavan, A., A. M. Bowen, A. Fernandez, A. J. Fielding, D. Kaminski, F. Moro, C. A. Muryn, M. D. Wise, A. Ruggi, E. J. L. McInnes, K. Severin, G. A. Timco, C. R. Timmel, F. Tuna, G. F. S. Whitehead, and R. E. P. Winpenny. 2015. “Engineering Coherent Interactions in Molecular Nanomagnet Dimers.” npj Quantum Information 1(1):15012. doi.org/10.1038/npjqi.2015.12.

Ardavan, A., O. Rival, J. J. L. Morton, S. J. Blundell, A. M. Tyryshkin, G. A. Timco, and R. E. P. Winpenny. 2007. “Will Spin-Relaxation Times in Molecular Magnets Permit Quantum Information Processing?” Physical Review Letters 98(5):057201. doi.org/10.1103/PhysRevLett.98.057201.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Ariciu, A.-M., D. H. Woen, D. N. Huh, L. E. Nodaraki, A. K. Kostopoulos, C. A. P. Goodwin, N. F. Chilton, E. J. L. McInnes, R. E. P. Winpenny, W. J. Evans, and F. Tuna. 2019. “Engineering Electronic Structure to Prolong Relaxation Times in Molecular Qubits by Minimising Orbital Angular Momentum.” Nature Communications 10(1):3330. doi.org/10.1038/s41467-019-11309-3.

Aslam, N., M. Pfender, P. Neumann, R. Reuter, A. Zappe, F. Fávaro de Oliveira, A. Denisenko, H. Sumiya, S. Onoda, J. Isoya, and J. Wrachtrup. 2017. “Nanoscale Nuclear Magnetic Resonance with Chemical Resolution.” Science 357(6346):67–71. doi:10.1126/science.aam8697.

Atzori, M., A. Chiesa, E. Morra, M. Chiesa, L. Sorace, S. Carretta, and R. Sessoli. 2018. “A Two-Qubit Molecular Architecture for Electron-Mediated Nuclear Quantum Simulation.” Chemical Science 9(29):6183–6192. doi.org/10.1039/C8SC01695J.

Atzori, M., L. Tesi, E. Morra, M. Chiesa, L. Sorace, and R. Sessoli. 2016. “Room-Temperature Quantum Coherence and Rabi Oscillations in Vanadyl Phthalocyanine: Toward Multifunctional Molecular Spin Qubits.” Journal of the American Chemical Society 138(7):2154–2157. doi.org/10.1021/jacs.5b13408.

Augenbraun, B. L., Z. D. Lasner, A. Frenett, H. Sawaoka, C. Miller, T. C. Steimle, and J. M. Doyle. 2020. “Laser-cooled Polyatomic Molecules for Improved Electron Electric Dipole Moment Searches.” New Journal of Physics 22(2):022003. doi.org/10.1088/1367-2630/ab687b.

Awschalom, D. D., R. Hanson, J. Wrachtrup, and B. B. Zhou. 2018. “Quantum Technologies with Optically Interfaced Solid-State Spins.” Nature Photonics 12(9):516–527. doi.org/10.1038/s41566-018-0232-2.

Bader, K., D. Dengler, S. Lenz, B. Endeward, S.-D. Jiang, P. Neugebauer, and J. van Slageren. 2014. “Room Temperature Quantum Coherence in a Potential Molecular Qubit.” Nature Communications 5(1):5304. doi.org/10.1038/ncomms6304.

Bader, K., S. H. Schlindwein, D. Gudat, and J. van Slageren. 2017. “Molecular Qubits Based on Potentially Nuclear-Spin-Free Nickel Ions.” Physical Chemistry Chemical Physics 19(3):2525–2529. doi.org/10.1039/C6CP08161D.

Baek, S. H., F. Borsa, Y. Furukawa, Y. Hatanaka, S. Kawakami, K. Kumagai, B. J. Suh, and A. Cornia. 2005. “57Fe NMR and Relaxation by Strong Collision in the Tunneling Regime in the Molecular Nanomagnet Fe8.” Physical Review B 71(21):214436. doi.org/10.1103/PhysRevB.71.214436.

Baker, M. L., S. J. Blundell, N. Domingo, and S. Hill. 2015. “Spectroscopy Methods for Molecular Nanomagnets.” In Molecular Nanomagnets and Related Phenomena, edited by S. Gao, 231–291. Berlin: Springer Berlin Heidelberg.

Baker, M. L., T. Lancaster, A. Chiesa, G. Amoretti, P. J. Baker, C. Barker, S. J. Blundell, S. Carretta, D. Collison, H. U. Güdel, T. Guidi, E. J. L. McInnes, J. S. Möller, H. Mutka, J. Ollivier, F. L. Pratt, P. Santini, F. Tuna, P. L. W. Tregenna-Piggott, I. J. Vitorica-Yrezabal, G. A. Timco, and R. E. P. Winpenny. 2016. “Studies of a Large Odd-Numbered Odd-Electron Metal Ring: Inelastic Neutron Scattering and Muon Spin Relaxation Spectroscopy of Cr8Mn.” Chemistry: A European Journal 22(5):1779–1788. doi.org/10.1002/chem.201503431.

Bao, F., H. Deng, D. Ding, R. Gao, X. Gao, C. Huang, X. Jiang, H-S. Ku, Z. Li, X. Ma, X. Ni, J. Qin, Z. Song, H. Sun, C. Tang, T. Wang, F. Wu, T. Xia, W. Yu, F. Zhang, G. Zhang, X. Zhang, J. Zhou, X. Zhu, Y. Shi, J. Chen, H-H Zhao, and C. Deng. 2022. “Fluxonium: An Alternative Qubit Platform for High-Fidelity Operations.” Physical Review Letters 129(1):010502. doi.org/10.1103/PhysRevLett.129.010502.

Barclay, P. E., K.-M. C. Fu, C. Santori, A. Faraon, and R. G. Beausoleil. 2011. “Hybrid Nanocavity Resonant Enhancement of Color Center Emission in Diamond.” Physical Review X 1(1):011007. doi.org/10.1103/PhysRevX.1.011007.

Barnes, A. B., G. De Paëpe, P. C. A. van der Wel, K. N. Hu, C. G. Joo, V. S. Bajaj, M. L. Mak-Jurkauskas, J. R. Sirigiri, J. Herzfeld, R. J. Temkin, and R. G. Griffin. 2008. “High-Field Dynamic Nuclear Polarization for Solid and Solution Biological NMR.” Applied Magnetic Resonance 34(3):237–263. doi.org/10.1007/s00723-008-0129-1.

Barra, A.-L., A. Caneschi, A. Cornia, D. Gatteschi, L. Gorini, L.-P. Heiniger, R. Sessoli, and L. Sorace. 2007. “The Origin of Transverse Anisotropy in Axially Symmetric Single Molecule Magnets.” Journal of the American Chemical Society 129(35):10754–10762. doi.org/10.1021/ja0717921.

Barra, A. L., D. Gatteschi, and R. Sessoli. 1997. “High-frequency EPR Spectra of a Molecular Nanomagnet: Understanding Quantum Tunneling of the Magnetization.” Physical Review B 56(13):8192–8198. doi.org/10.1103/PhysRevB.56.8192.

Barry, J. F., D. J. McCarron, E. B. Norrgard, M. H. Steinecker, and D. DeMille. 2014. “Magneto-Optical Trapping of a Diatomic Molecule.” Nature 512(7514):286–289. doi.org/10.1038/nature13634.

Baumann, S., W. Paul, T. Choi, C. P. Lutz, A. Ardavan, and A. J. Heinrich. 2015. “Electron Paramagnetic Resonance of Individual Atoms on a Surface.” Science 350(6259):417–420. doi.org/10.1126/science.aac8703.

Bayliss, S. L., D. W. Laorenza, P. J. Mintun, B. D. Kovos, D. E. Freedman, and D. D. Awschalom. 2020. “Optically Addressable Molecular Spins for Quantum Information Processing.” Science 370(6522):1309–1312. doi.org/10.1126/science.abb9352.

Bertaina, S., S. Gambarelli, T. Mitra, B. Tsukerblat, A. Müller, and B. Barbara. 2008. “Quantum Oscillations in a Molecular Magnet.” Nature 453(7192):203–206. doi.org/10.1038/nature06962.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Blackaby, W. J. M., K. L. M. Harriman, S. M. Greer, A. Folli, S. Hill, V. Krewald, M. F. Mahon, D. M. Murphy, M. Murugesu, E. Richards, E. Suturina, and M. K. Whittlesey. 2022. “Extreme g-Tensor Anisotropy and Its Insensitivity to Structural Distortions in a Family of Linear Two-Coordinate Ni(I) Bis-N-heterocyclic Carbene Complexes.” Inorganic Chemistry 61(3):1308–1315. doi.org/10.1021/acs.inorgchem.1c02413.

Bluhm, H., S. Foletti, D. Mahalu, V. Umansky, and A. Yacoby. 2010. “Enhancing the Coherence of a Spin Qubit by Operating It as a Feedback Loop That Controls Its Nuclear Spin Bath.” Physical Review Letters 105(21):216803. doi.org/10.1103/PhysRevLett.105.216803.

Bonizzoni, C., A. Ghirri, and M. Affronte. 2018. “Coherent Coupling of Molecular Spins with Microwave Photons in Planar Superconducting Resonators.” Advances in Physics: X 3(1):1435305. doi.org/10.1080/23746149.2018.1435305.

Boudalis, A. K., J. Robert, and P. Turek. 2018. “First Demonstration of Magnetoelectric Coupling in a Polynuclear Molecular Nanomagnet: Single-Crystal EPR Studies of [Fe3O(O2CPh)6(py)3]ClO4·py Under Static Electric Fields.” Chemistry: A European Journal 24(56):14896–14900. doi.org/10.1002/chem.201803038.

Burdick, R. K., G. C. Schatz, and T. Goodson, III. 2021. “Enhancing Entangled Two-Photon Absorption for Picosecond Quantum Spectroscopy.” Journal of the American Chemical Society 143(41):16930–16934. doi.org/10.1021/jacs.1c09728.

Bustamante, C. J., Y. R. Chemla, S. Liu, and M. D. Wang. 2021. “Optical Tweezers in Single-Molecule Biophysics.” Nature Reviews Methods Primers 1:25. doi.org/10.1038/s43586-021-00021-6.

Can, T. V., Q. Z. Ni, and R. G. Griffin. 2015. “Mechanisms of Dynamic Nuclear Polarization in Insulating Solids.” Journal of Magnetic Resonance 253:23–35. doi.org/10.1016/j.jmr.2015.02.005.

Canarie, E. R., S. M. Jahn, and S. Stoll. 2020. “Quantitative Structure-Based Prediction of Electron Spin Decoherence in Organic Radicals.” Physical Review Letters 11(9):3396–3400. doi.org/10.1021/acs.jpclett.0c00768.

Casacio, C. A., L. S. Madsen, A. Terrasson, M. Waleed, K. Barnscheidt, B. Hage, M. A. Taylor, and W. P. Bowen. 2021. “Quantum-Enhanced Nonlinear Microscopy.” Nature 594(7862):201–206. doi.org/10.1038/s41586-021-03528-w.

Chakov, N. E., S.-C. Lee, A. G. Harter, P. L. Kuhns, A. P. Reyes, S. O. Hill, N. S. Dalal, W. Wernsdorfer, K. A. Abboud, and G. Christou. 2006. “The Properties of the [Mn12O12(O2CR)16(H2O)4] Single-Molecule Magnets in Truly Axial Symmetry: [Mn12O12(O2CCH2Br)16(H2O)4]· 4CH2Cl2.” Journal of the American Chemical Society 128(21):6975–6989. doi.org/10.1021/ja060796n.

Chen, F., and Mukamel S. 2021. “Vibrational Hyper-Raman Molecular Spectroscopy with Entangled Photons.” ACS Photonics 8(9):2722–27. https://doi.org/10.1021/acsphotonics.1c00777.

Chen, J., C. Hu, J. F. Stanton, S. Hill, H.-P. Cheng, and X.-G. Zhang. 2020. “Decoherence in Molecular Electron Spin Qubits: Insights from Quantum Many-Body Simulations.” Physical Review Letters 11(6):2074–2078. doi.org/10.1021/Acs.Jpclett.0c00193.

Chen, M., C. Meng, Q. Zhang, C. Duan, F. Shi, and J. Du. 2017. “Quantum Metrology with Single Spins in Diamond Under Ambient Conditions.” National Science Review 5(3):346–355. doi.org/10.1093/Nsr/Nwx121.

Chen, Y., Y. Bae, and A. J. Heinrich. 2022. “Harnessing the Quantum Behavior of Spins on Surfaces.” Advanced Materials. doi.org/10.1002/Adma.202107534.

Chenu, A., and G. D. Scholes. 2015. “Coherence in Energy Transfer and Photosynthesis.” Annual Review of Physical Chemistry 66(1):69–96. doi.org/10.1146/Annurev-Physchem-040214-121713.

Cheuk, L. W., L. Anderegg, B. L. Augenbraun, Y. Bao, S. Burchesky, W. Ketterle, and J. M. Doyle. 2018. “Λ-Enhanced Imaging of Molecules in an Optical Trap.” Physical Review Letters 121(8):083201. doi.org/10.1103/PhysRevLett.121.083201.

Chiesa, A., T. Guidi, S. Carretta, S. Ansbro, G. A. Timco, I. Vitorica-Yrezabal, E. Garlatti, G. Amoretti, R. E. P. Winpenny, and P. Santini. 2017. “Magnetic Exchange Interactions in the Molecular Nanomagnet Mn12Physical Review Letters 119(21):217202. doi.org/10.1103/Physrevlett.119.217202.

Chiesa, A., E. Macaluso, F. Petiziol, S. Wimberger, P. Santini, and S. Carretta. 2020. “Molecular Nanomagnets as Qubits with Embedded Quantum-Error Correction.” Journal of Physical Chemistry Letters 11(20):8610–8615. doi.org/10.1021/acs.jpclett.0c02213.

Choi, T. 2019. “Studies of Single Atom Magnets Via Scanning Tunneling Microscopy.” Journal of Magnetism and Magnetic Materials 481:150–155. doi.org/10.1016/J.Jmmm.2019.03.007.

Chou, C.-W., C. Kurz, D. B. Hume, P. N. Plessow, D. R. Leibrandt, and D. Leibfried. 2017. “Preparation and Coherent Manipulation of Pure Quantum States of a Single Molecular Ion.” Nature 545(7653):203–207. doi.org/10.1038/Nature22338.

Cimini, V., M. Mellini, G. Rampioni, M. Sbroscia, L. Leoni, M. Barbieri, and I. Gianani. 2019. “Adaptive Tracking of Enzymatic Reactions with Quantum Light.” Optics Express 27(24):35245. doi.org/10.1364/oe.27.035245.

Cini, A., M. Mannini, F. Totti, M. Fittipaldi, G. Spina, A. Chumakov, R. Rüffer, A. Cornia, and R. Sessoli. 2018. “Mössbauer Spectroscopy of a Monolayer of Single Molecule Magnets.” Nature Communications 9(1):480. doi.org/10.1038/S41467-018-02840-W.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Collett, C. A., K.-I. Ellers, N. Russo, K. R. Kittilstved, G. A. Timco, R. E. P. Winpenny, and J. R. Friedman. 2019. “A Clock Transition in the Cr7Mn Molecular Nanomagnet.” Magnetochemistry 5(1):4. doi.org/10.3390/magnetochemistry5010004.

Collett, C. A., P. Santini, S. Carretta, and J. R. Friedman. 2020. “Constructing Clock-Transition-Based Two-Qubit Gates from Dimers of Molecular Nanomagnets.” Physical Review Research 2(3):032037. doi.org/10.1103/Physrevresearch.2.032037.

Collopy, A. L., S. Ding, Y. Wu, I. A. Finneran, L. Anderegg, B. L. Augenbraun, J. M. Doyle, and J. Ye. 2018. “3D Magneto-Optical Trap of Yttrium Monoxide.” Physical Review Letters 121(21):213201. doi.org/10.1103/Physrevlett.121.213201.

Cruickshank, P. A. S., D. R. Bolton, D. A. Robertson, R. I. Hunter, R. J. Wylde, and G. M. Smith. 2009. “A Kilowatt Pulsed 94 Ghz Electron Paramagnetic Resonance Spectrometer with High Concentration Sensitivity, High Instantaneous Bandwidth, and Low Dead Time.” Review of Scientific Instruments 80(10):103102. doi.org/10.1063/1.3239402.

Defienne, H., B. Ndagano, A. Lyons, and D. Faccio. 2021. “Polarization Entanglement-Enabled Quantum Holography.” Nature Physics 17(5):591–597. doi.org/10.1038/S41567-020-01156-1.

Degen, C. L., M. Poggio, H. J. Mamin, C. T. Rettner, and D. Rugar. 2009. “Nanoscale Magnetic Resonance Imaging.” Proceedings of the National Academy of Sciences USA 106(5):1313–1317. doi.org/10.1073/Pnas.0812068106.

del Barco, E., A. D. Kent, S. Hill, J. M. North, N. S. Dalal, E. M. Rumberger, D. N. Hendrickson, N. Chakov, and G. Christou. 2005. “Magnetic Quantum Tunneling in the Single-Molecule Magnet Mn12-Acetate.” Journal of Low Temperature Physics 140(1):119–174. doi.org/10.1007/s10909-005-6016-3.

Di Rosa, M. D. 2004. “Laser-Cooling Molecules.” European Physical Journal D—Atomic, Molecular, Optical, and Plasma Physics 31(2):395–402. doi.org/10.1140/Epjd/E2004-00167-2.

Dickerson, C. E., C. Chang, H. Guo, and A. N. Alexandrova. 2022. “Fully Saturated Hydrocarbons as Hosts of Optical Cycling Centers.” Journal of Physical Chemistry A 126(51):9644–9650. doi.org/10.1021/Acs.Jpca.2c06647.

Dickerson, C. E., H. Guo, A. J. Shin, B. L. Augenbraun, J. R. Caram, W. C. Campbell, and A. N. Alexandrova. 2021. “Franck-Condon Tuning of Optical Cycling Centers by Organic Functionalization.” Physical Review Letters 126(12):123002. doi.org/10.1103/Physrevlett.126.123002.

Dickerson, C. E., H. Guo, G.-Z. Zhu, E. R. Hudson, J. R. Caram, W. C. Campbell, and A. N. Alexandrova. 2021. “Optical Cycling Functionalization of Arenes.” Physical Review Letters 12(16):3989–3995. doi.org/10.1021/Acs.Jpclett.1c00733.

Dolde, F., I. Jakobi, B. Naydenov, N. Zhao, S. Pezzagna, C. Trautmann, J. Meijer, P. Neumann, F. Jelezko, and J. Wrachtrup. 2013. “Room-Temperature Entanglement Between Single Defect Spins in Diamond.” Nature Physics 9(3):139–143. doi.org/10.1038/Nphys2545.

Dorfman, K. E., F. Schlawin, and S. Mukamel. 2014. “Stimulated Raman Spectroscopy with Entangled Light: Enhanced Resolution and Pathway Selection.” Physical Review Letters 5(16):2843–2849. doi.org/10.1021/Jz501124a.

Dressel, M., B. Gorshunov, K. Rajagopal, S. Vongtragool, and A. A. Mukhin. 2003. “Quantum Tunneling and Relaxation in Mn12-Acetate Studied by Magnetic Spectroscopy.” Physical Review B 67(6):060405. doi.org/10.1103/Physrevb.67.060405.

Du, M., R. F. Ribeiro, and J. Yuen-Zhou. 2019. “Remote Control of Chemistry in Optical Cavities.” Chem 5(5):1167–1181. doi.org/10.1016/J.Chempr.2019.02.009.

Duan, L. M., M. D. Lukin, J. I. Cirac, and P. Zoller. 2001. “Long-Distance Quantum Communication with Atomic Ensembles and Linear Optics.” Nature 414(6862):413–418. doi.org/10.1038/35106500.

Duan, R., J. N. Mastron, Y. Song, and K. J. Kubarych. 2021. “Isolating Polaritonic 2D-IR Transmission Spectra.” Journal of Physical Chemistry Letters 12(46):11406–11414. doi.org/10.1021/acs.jpclett.1c03198.

Durbin, S. M. 2022. “Proposal for Entangled X-Ray Beams.” Journal of Applied Physics 131(22):224401. doi.org/10.1063/5.0091947.

Durkan, C., and M. E. Welland. 2002. “Electronic Spin Detection in Molecules Using Scanning-Tunneling-Microscopy-Assisted Electron-Spin Resonance.” Applied Physics Letters 80(3):458–460. doi.org/10.1063/1.1434301.

Ebadi, S., T. T. Wang, H. Levine, A. Keesling, G. Semeghini, A. Omran, D. Bluvstein, R. Samajdar, H. Pichler, W. W. Ho, S. Choi, S. Sachdev, M. Greiner, V. Vuletić, and M. D. Lukin. 2021. “Quantum Phases of Matter on a 256-Atom Programmable Quantum Simulator.” Nature 595:227–232. doi.org/10.1038/s41586-021-03582-4.

Ebbesen, T. W. 2016. “Hybrid Light–Matter States in a Molecular and Material Science Perspective.” Accounts of Chemical Research 49(11):2403–2412. doi.org/10.1021/Acs.Accounts.6b00295.

Eddins, A. W., C. C. Beedle, D. N. Hendrickson, and J. R. Friedman. 2014. “Collective Coupling of a Macroscopic Number of Single-Molecule Magnets with a Microwave Cavity Mode.” Physical Review Letters 112(12):120501. doi.org/10.1103/Physrevlett.112.120501.

Engel, G. S., T. R. Calhoun, E. L. Read, T.-K. Ahn, T. Mančal, Y.-C. Cheng, R. E. Blankenship, and G. R. Fleming. 2007. “Evidence for Wavelike Energy Transfer Through Quantum Coherence in Photosynthetic Systems.” Nature 446(7137):782–786. doi.org/10.1038/Nature05678.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Eshun, A., B. Gu, O. Varnavski, S. Asban, K. E. Dorfman, S. Mukamel, and T. Goodson, III. 2021. “Investigations of Molecular Optical Properties Using Quantum Light and Hong–Ou–Mandel Interferometry.” Journal of the American Chemical Society 143(24):9070–9081. doi.org/10.1021/Jacs.1c02514.

Eshun, A., O. Varnavski, J. P. Villabona-Monsalve, R. K. Burdick, and T. Goodson, III. 2022. “Entangled Photon Spectroscopy.” Accounts of Chemical Research 55(7):991–1003. doi.org/10.1021/Acs.Accounts.1c00687.

Evans, R. E., M. K. Bhaskar, D. D. Sukachev, C. T. Nguyen, A. Sipahigil, M. J. Burek, B. Machielse, G. H. Zhang, A. S. Zibrov, E. Bielejec, H. Park, M. Lončar, and M. D. Lukin. 2018. “Photon-Mediated Interactions Between Quantum Emitters in a Diamond Nanocavity.” Science 362(6415):662–665. doi.org/10.1126/Science.Aau4691.

Farina, M., and J. C. M. Hwang. 2020. “Scanning Microwave Microscopy for Biological Applications: Introducing the State of the Art and Inverted SMM.” IEEE Microwave Magazine 21(10):52–59. doi.org/10.1109/MMM.2020.3008239.

Fataftah, M. S., S. L. Bayliss, D. W. Laorenza, X. Wang, B. T. Phelan, C. B. Wilson, P. J. Mintun, B. D. Kovos, M. R. Wasielewski, S. Han, M. S. Sherwin, D. D. Awschalom, and D. E. Freedman. 2020. “Trigonal Bipyramidal V3+ Complex as an Optically Addressable Molecular Qubit Candidate.” Journal of the American Chemical Society 142(48):20400–20408. doi.org/10.1021/Jacs.0c08986.

Fataftah, M. S., J. M. Zadrozny, S. C. Coste, M. J. Graham, D. M. Rogers, and D. E. Freedman. 2016. “Employing Forbidden Transitions as Qubits in a Nuclear Spin-Free Chromium Complex.” Journal of the American Chemical Society 138(4):1344–1348. doi.org/10.1021/Jacs.5b11802.

Fei, H-B., B. M. Jost, S. Popescu, B. E. A. Saleh, and M. C. Teich. 1997. “Entanglement-Induced Two-Photon Transparency.” Physical Review Letters 78(9):1679–1682. doi.org/10.1103/PhysRevLett.78.1679.

Feng, L., K.-Y. Wang, J. Willman, and H.-C. Zhou. 2020. “Hieracrchy in Metal-Organic Frameworks” Journal of the American Chemical Society 6(3):359–367. doi.org/10.1021/acscentsci.0c00158

Feng, X., J. Liu, T. D. Harris, S. Hill, and J. R. Long. 2012. “Slow Magnetic Relaxation Induced by a Large Transverse Zero-Field Splitting in a MnIIreIV(CN)2 Single-Chain Magnet.” Journal of the American Chemical Society 134(17):7521–7529. doi.org/10.1021/Ja301338d.

Ferrando-Soria, J., S. A. Magee, A. Chiesa, S. Carretta, P. Santini, I. J. Vitorica-Yrezabal, F. Tuna, G. F. S. Whitehead, S. Sproules, K. M. Lancaster, A.-L. Barra, G. A. Timco, E. J. L. Mcinnes, and R. E. P. Winpenny. 2016. “Switchable Interaction in Molecular Double Qubits.” Chem 1(5):727–752. doi.org/10.1016/J.Chempr.2016.10.001.

Ferrando-Soria, J., E. M. Pineda, A. Chiesa, A. Fernandez, S. A. Magee, S. Carretta, P. Santini, I. J. Vitorica-Yrezabal, F. Tuna, G. A. Timco, E. J. L. McInnes, and R. E. P. Winpenny. 2016. “A Modular Design of Molecular Qubits to Implement Universal Quantum Gates.” Nature Communications 7(1):11377. doi.org/10.1038/ncomms11377.

Fittipaldi, M., A. Cini, G. Annino, A. Vindigni, A. Caneschi, and R. Sessoli. 2019. “Electric Field Modulation of Magnetic Exchange in Molecular Helices.” Nature Materials 18(4):329–334. doi.org/10.1038/S41563-019-0288-5.

Foletti, S., H. Bluhm, D. Mahalu, V. Umansky, and A. Yacoby. 2009. “Universal Quantum Control of Two-Electron Spin Quantum Bits Using Dynamic Nuclear Polarization.” Nature Physics 5(12):903–908. doi.org/10.1038/Nphys1424.

Fortman, B., J. Pena, K. Holczer, and S. Takahashi. 2020. “Demonstration of NV-Detected ESR Spectroscopy at 115 Ghz and 4.2 T.” Applied Physics Letters 116(17):174004. doi.org/10.1063/5.0006014.

Fröhlich, H. 1968. “Long-Range Coherence and Energy Storage in Biological Systems.” International Journal of Quantum Chemistry 2(5):641–649. doi.org/10.1002/Qua.560020505.

Fu, P.-X., S. Zhou, Z. Liu, C.-H. Wu, Y.-H. Fang, Z.-R. Wu, X.-Q. Tao, J.-Y. Yuan, Y.-X. Wang, S. Gao, and S.-D. Jiang. 2022. “Multiprocessing Quantum Computing Through Hyperfine Couplings in Endohedral Fullerene Derivatives.” Angewandte Chemie International Edition 61(52):E202212939. doi.org/10.1002/Anie.202212939.

Fukuzawa, H., and K. Ueda. 2020. “X-Ray Induced Ultrafast Dynamics in Atoms, Molecules, and Clusters: Experimental Studies at an X-Ray Free-Electron Laser Facility SACLA and Modelling.” Advances in Physics: X 5(1):1785327. doi.org/10.1080/23746149.2020.1785327.

Furue, S., T. Kohmoto, M. Kunitomo, and Y. Fukuda. 2005. “Optical Induction of Magnetization and Observation of Fast Spin Dynamics in Aqueous Solutions of Copper Ions.” Physics Letters A 345(4):415–422. doi.org/10.1016/J.Physleta.2005.07.028.

Furukawa, Y., K. Watanabe, K. Kumagai, F. Borsa, and D. Gatteschi. 2001. “Magnetic Structure and Spin Dynamics of the Ground State of the Molecular Cluster Mn12O12 Acetate Studied by 55Mn NMR.” Physical Review B 64(10):104401. doi.org/10.1103/Physrevb.64.104401.

Gaita-Ariño, A., F. Luis, S. Hill, and E. Coronado. 2019. “Molecular Spins for Quantum Computation.” Nature Chemistry 11(4):301–309. doi.org/10.1038/S41557-019-0232-Y.

Gallez, B. 2022. “Interview with Professor Graham Smith on the Occasion of His Bruker Prize 2022.” EPR Newsletter.

Gan, Z., I. Hung, X. Wang, J. Paulino, G. Wu, I. M. Litvak, P. L. Gor’kov, W. W. Brey, P. Lendi, J. L. Schiano, M. D. Bird, I. R. Dixon, J. Toth, G. S. Boebinger, and T. A. Cross. 2017. “NMR Spectroscopy Up to 35.2T Using a Series-Connected Hybrid Magnet.” Journal of Magnetic Resonance 284:125–136. doi.org/10.1016/J.Jmr.2017.08.007.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Garlatti, E., A. Chiesa, T. Guidi, G. Amoretti, P. Santini, and S. Carretta. 2019. “Unravelling the Spin Dynamics of Molecular Nanomagnets with Four-Dimensional Inelastic Neutron Scattering.” European Journal of Inorganic Chemistry 2019 (8):1106–1118. doi.org/10.1002/Ejic.201801050.

Garlatti, E., T. Guidi, S. Ansbro, P. Santini, G. Amoretti, J. Ollivier, H. Mutka, G. Timco, I. J. Vitorica-Yrezabal, G. F. S. Whitehead, R. E. P. Winpenny, and S. Carretta. 2017. “Portraying Entanglement Between Molecular Qubits with Four-Dimensional Inelastic Neutron Scattering.” Nature Communications 8(1):14543. doi.org/10.1038/Ncomms14543.

Garlatti, E., L. Tesi, A. Lunghi, M. Atzori, D. J. Voneshen, P. Santini, S. Sanvito, T. Guidi, R. Sessoli, and S. Carretta. 2020. “Unveiling Phonons in a Molecular Qubit with Four-Dimensional Inelastic Neutron Scattering and Density Functional Theory.” Nature Communications 11(1):1751. doi.org/10.1038/S41467-020-15475-7.

Gatteschi, D., A. L. Barra, A. Caneschi, A. Cornia, R. Sessoli, and L. Sorace. 2006. “EPR of Molecular Nanomagnets.” Coordination Chemistry Reviews 250(11):1514–1529. doi.org/10.1016/J.Ccr.2006.02.006.

Gehring, P., J. M. Thijssen, and H. S. J. van der Zant. 2019. “Single-Molecule Quantum-Transport Phenomena in Break Junctions.” Nature Reviews Physics 1(6):381–396. doi.org/10.1038/S42254-019-0055-1.

George, J., A. Shalabney, J.A. Hutchison, C. Genet, and T.W. Ebbesen. 2015. “Liquid-Phase Vibrational Strong Coupling.” Journal of Physical Chemistry Letters 6(6):1027–1031. doi.org/10.1021/acs.jpclett.5b00204.

Georgiades, N. Ph., E. S. Polzik, K. Edamatsu, H. J. Kimble, and A. S. Parkins. 1995. “Nonclassical Excitation for Atoms in a Squeezed Vacuum.” Physical Review Letters 75 (19):3426–3429. doi.org/10.1103/Physrevlett.75.3426.

Ghosh, T., J. Marbey, W. Wernsdorfer, S. Hill, K. A. Abboud, and G. Christou. 2021. “Exchange-Biased Quantum Tunnelling of Magnetization in a [Mn3]2 Dimer of Single-Molecule Magnets with Rare Ferromagnetic Inter-Mn3 Coupling.” Physical Chemistry Chemical Physics 23(14):8854–8867. doi.org/10.1039/D0CP06611G.

Giansiracusa, M. J., E. Moreno-Pineda, R. Hussain, R. Marx, M. Martínez Prada, P. Neugebauer, S. Al-Badran, D. Collison, F. Tuna, J. Van Slageren, S. Carretta, T. Guidi, E. J. L. Mcinnes, R. E. P. Winpenny, and N. F. Chilton. 2018. “Measurement of Magnetic Exchange in Asymmetric Lanthanide Dimetallics: Toward a Transferable Theoretical Framework.” Journal of the American Chemical Society 140(7):2504–2513. doi.org/10.1021/Jacs.7b10714.

Gimeno, I., W. Kersten, M. C. Pallarés, P. Hermosilla, M. J. Martínez-Pérez, M. D. Jenkins, A. Angerer, C. Sánchez-Azqueta, D. Zueco, J. Majer, A. Lostao, and F. Luis. 2020. “Enhanced Molecular Spin-Photon Coupling at Superconducting Nano-constrictions.” ACS Nano 14(7):8707–8715. doi.org/10.1021/Acsnano.0c03167.

Gimeno, I., A. Urtizberea, J. Román-Roche, D. Zueco, A. Camón, P. J. Alonso, O. Roubeau, and F. Luis. 2021. “Broad-Band Spectroscopy of a Vanadyl Porphyrin: A Model Electronuclear Spin Qudit.” Chemical Science 12(15):5621–5630. doi.org/10.1039/D1SC00564B.

Giovannetti, V. 2004. “Quantum-Enhanced Measurements: Beating the Standard Quantum Limit.” Science 306(5700):1330–1336. https://doi.org/10.1126/science.1104149.

Giri, S. K., and G. C. Schatz. 2022. “Manipulating Two-Photon Absorption of Molecules Through Efficient Optimization of Entangled Light.” Physical Review Letters 13(43):10140–10146. doi.org/10.1021/Acs.Jpclett.2c02842.

Godejohann, F., A. V. Scherbakov, S. M. Kukhtaruk, A. N. Poddubny, D. D. Yaremkevich, M. Wang, A. Nadzeyka, D. R. Yakovlev, A. W. Rushforth, A. V. Akimov, and M. Bayer. 2020. “Magnon Polaron Formed by Selectively Coupled Coherent Magnon and Phonon Modes of a Surface Patterned Ferromagnet.” Physical Review B 102(14):144438. doi.org/10.1103/Physrevb.102.144438.

Godfrin, C., A. Ferhat, R. Ballou, S. Klyatskaya, M. Ruben, W. Wernsdorfer, and F. Balestro. 2017. “Operating Quantum States in Single Magnetic Molecules: Implementation of Grover’s Quantum Algorithm.” Physical Review Letters 119(18):187702. doi.org/10.1103/Physrevlett.119.187702.

Godfrin, C., S. Thiele, A. Ferhat, S. Klyatskaya, M. Ruben, W. Wernsdorfer, and F. Balestro. 2017. “Electrical Read-Out of a Single Spin Using an Exchange-Coupled Quantum Dot.” ACS Nano 11(4):3984–3989. doi.org/10.1021/Acsnano.7b00451.

Goldfarb, D. 2017. “ELDOR-Detected NMR.” eMagRes 563:101–114.

Gould, C. A., J. Marbey, V. Vieru, D. A. Marchiori, R. D. Britt, L. F. Chibotaru, S. Hill, and J. R. Long. 2021. “Isolation of a Triplet Benzene Dianion.” Nature Chemistry 13(10):1001–1005. doi.org/10.1038/S41557-021-00737-8.

Graham, M. J., J. M. Zadrozny, M. Shiddiq, J. S. Anderson, M. S. Fataftah, S. Hill, and D. E. Freedman. 2014. “Influence of Electronic Spin and Spin–Orbit Coupling on Decoherence in Mononuclear Transition Metal Complexes.” Journal of the American Chemical Society 136(21):7623–7626. doi.org/10.1021/Ja5037397.

Greer, S. M., J. Mckay, K. M. Gramigna, C. M. Thomas, S. A. Stoian, and S. Hill. 2018. “Probing Fe–V Bonding in a C3-Symmetric Heterobimetallic Complex.” Inorganic Chemistry 57(10):5870–5878. doi.org/10.1021/Acs.Inorgchem.8b00280.

Gross, L., K.-H. Rieder, F. Moresco, S. M. Stojkovic, A. Gourdon, and C. Joachim. 2005. “Trapping and Moving Metal Atoms with a Six-Leg Molecule.” Nature Materials 4(12):892–895. doi.org/10.1038/Nmat1529.

Gu, B., and I. Franco. 2018. “Optical Absorption Properties of Laser-Driven Matter.” Physical Review A 98(6):063412. doi.org/10.1103/Physreva.98.063412.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Gu, B., and S. Mukamel. 2020. “Manipulating Two-Photon-Absorption of Cavity Polaritons by Entangled Light.” Journal of Physical Chemistry Letters 11(19):8177–8182. doi.org/10.1021/Acs.Jpclett.0c02282.

Gu, B., and S. Mukamel. 2021. “Optical-Cavity Manipulation of Conical Intersections and Singlet Fission in Pentacene Dimers.” Journal of Physical Chemistry Letters 12(8):2052–2056. doi.org/10.1021/acs.jpclett.0c03829.

Gu, B., D. Keefer, F. Aleotti, A. Nenov, M. Garavelli, and S. Mukamel. 2021. “Photoisomerization Transition State Manipulation by Entangled Two-Photon Absorption.” Proceedings of the National Academy of Sciences USA 118(47):E2116868118. doi.org/10.1073/Pnas.2116868118.

Guo, H., C. E. Dickerson, A. J. Shin, C. Zhao, T. L. Atallah, J. R. Caram, W. C. Campbell, and A. N. Alexandrova. 2021. “Surface Chemical Trapping of Optical Cycling Centers.” Physical Chemistry Chemical Physics 23(1):211–218. doi.org/10.1039/D0CP04525J.

Guzman, A. R., M. R. Harpham, Ö. Süzer, M. M. Haley, and T. G. Goodson, III. 2010. “Spatial Control of Entangled Two-Photon Absorption with Organic Chromophores.” Journal of the American Chemical Society 132(23):7840–7841. doi.org/10.1021/Ja1016816.

Hallas, C., N. B. Vilas, L. Anderegg, P. Robichaud, A. Winnicki, C. Zhang, L. Cheng, and J. M. Doyle. 2023. “Optical Trapping of a Polyatomic Molecule in an l-Type Parity Doublet State.” Physical Review Letters 130:153202. doi.org/10.1103/PhysRevLett.130.153202.

Hameroff, S., and R. Penrose. 2014. “Consciousness in the Universe: A Review of the ‘Orch OR’ Theory.” Physics of Life Reviews 11(1):39–78. doi.org/10.1016/J.Plrev.2013.08.002.

Harding, R. T., S. Zhou, J. Zhou, T. Lindvall, W. K. Myers, A. Ardavan, G. A. D. Briggs, K. Porfyrakis, and E. A. Laird. 2017. “Spin Resonance Clock Transition of the Endohedral Fullerene 15N@C60.” Physical Review Letters 119(14):140801. doi.org/10.1103/Physrevlett.119.140801.

Harmer, J. R. 2016. “Hyperfine Spectroscopy–ENDOR.” eMagRes 5:1493–1514.

Harpham, M. R., Ö. Süzer, C.-Q. Ma, P. Bäuerle, and T. Goodson, III. 2009. “Thiophene Dendrimers as Entangled Photon Sensor Materials.” Journal of the American Chemical Society 131(3):973–979. doi.org/10.1021/Ja803268s.

Harvey, S. M., and M. R. Wasielewski. 2021. “Photogenerated Spin-Correlated Radical Pairs: From Photosynthetic Energy Transduction to Quantum Information Science.” Journal of the American Chemical Society 143(38):15508–15529. doi.org/10.1021/Jacs.1c07706.

Hassan, A. K., L. A. Pardi, J. Krzystek, A. Sienkiewicz, P. Goy, M. Rohrer, and L. C. Brunel. 2000. “Ultrawide Band Multifrequency High-Field EMR Technique: A Methodology for Increasing Spectroscopic Information.” Journal of Magnetic Resonance 142(2):300–312. doi.org/10.1006/Jmre.1999.1952.

Hassan, M. A., T. Elrifai, A. Sakr, M. Kern, K. Lips, and J. Anders. 2021. “A 14-Channel 7 GHz VCO-Based EPR-on-a-Chip Sensor with Rapid Scan Capabilities.” 2021 IEEE Sensors, Virtual Conference, Oct. 31–Nov. 3, 2021.

Hassan, M. A., M. Kern, A. Chu, G. Kalra, E. Shabratova, A. Tsarapkin, N. Mackinnon, K. Lips, C. Teutloff, R. Bittl, J. G. Korvink, and J. Anders. 2022. “Towards Single-Cell Pulsed EPR Using VCO-Based EPR-on-a-Chip Detectors.” Frequenz 76(11–12):699–717. doi.org/10.1515/Freq-2022-0096.

Hay, M. A., A. Sarkar, G. A. Craig, L. Bhaskaran, J. Nehrkorn, M. Ozerov, K. E. R. Marriott, C. Wilson, G. Rajaraman, S. Hill, and M. Murrie. 2019. “In-Depth Investigation of Large Axial Magnetic Anisotropy in Monometallic 3d Complexes Using Frequency Domain Magnetic Resonance and Ab Initio Methods: A Study of Trigonal Bipyramidal Co(II).” Chemical Science 10(25):6354–6361. doi.org/10.1039/C9SC00987F.

Hecht, B., B. Sick, U. P. Wild, V. Deckert, R. Zenobi, O. J. F. Martin, and D. W. Pohl. 2000. “Scanning Near-Field Optical Microscopy with Aperture Probes: Fundamentals and Applications.” Journal of Chemical Physics 112(18):7761–7774. doi.org/10.1063/1.481382.

Heinrich, A. J., J. A. Gupta, C. P. Lutz, and D. M. Eigler. 2004. “Single-Atom Spin-Flip Spectroscopy.” Science 306(5695): 466–469. doi.org/10.1126/Science.1101077.

Heinz, T., O. Shpyrko, D. Basov, N. Berrah, P. Bucksbaum, T. Devereaux, D. Fritz, K. Gaffney, O. Gessner, V. Gopalan, Z. Hasan, A. Lanzara, T. Martinez, A. Millis, S. Mukamel, M. Murnane, K. Nelson, R. Prasankumar, D. Reis, K. Schafer, G. Scholes, Z. X. Shen, A. Stolow, H. Wen, M. Wolf, D. Xiao, L. Young, B. Garrett, L. Horton, H. Kerch, J. Krause, T. Settersten, L. Wilson, K. Runkles, T. Anderson, G. Chui, and E. Rutherford. 2017. “Basic Energy Sciences Roundtable: Opportunities for Basic Research at the Frontiers of XFEL Ultrafast Science.” Office of Scientific and Technical Information (OSTI). doi.org/10.2172/1616251.

Hemmer, P. 2013. “Toward Molecular-Scale MRI.” Science 339(6119):529–530. doi.org/10.1126/Science.1233222.

Hensen, B., H. Bernien, A. E. Dréau, A. Reiserer, N. Kalb, M. S. Blok, J. Ruitenberg, R. F. L. Vermeulen, R. N. Schouten, C. Abellán, W. Amaya, V. Pruneri, M. W. Mitchell, M. Markham, D. J. Twitchen, D. Elkouss, S. Wehner, T. H. Taminiau, and R. Hanson. 2015. “Loophole-Free Bell Inequality Violation Using Electron Spins Separated by 1.3 Kilometres.” Nature 526(7575):682–686. doi.org/10.1038/Nature15759.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Hernandez, J. M., X. X. Zhang, F. Luis, J. Tejada, J. R. Friedman, M. P. Sarachik, and R. Ziolo. 1997. “Evidence for Resonant Tunneling of Magnetization in Mn12 Acetate Complex.” Physical Review B 55(9):5858–5865. doi.org/10.1103/Physrevb.55.5858.

Higgins, J. S., M. A. Allodi, L. T. Lloyd, J. P. Otto, S. H. Sohail, R. G. Saer, R. E. Wood, S. C. Massey, P. C. Ting, R. E. Blankenship, and G. S. Engel. 2021. “Redox Conditions Correlated with Vibronic Coupling Modulate Quantum Beats in Photosynthetic Pigment-Protein Complexes.” Proceedings of the National Academy of Sciences USA 118(49). doi.org/10.1073/Pnas.2112817118.

Higgins, J. S., L. T. Lloyd, S. H. Sohail, M. A. Allodi, J. P. Otto, R. G. Saer, R. E. Wood, S. C. Massey, P.-C. Ting, R. E. Blankenship, and G. S. Engel. 2021. “Photosynthesis Tunes Quantum-Mechanical Mixing of Electronic and Vibrational States to Steer Exciton Energy Transfer.” Proceedings of the National Academy of Sciences USA 118(11):E2018240118. doi.org/10.1073/Pnas.2018240118.

Hill, S. 2013. “Magnetization Tunneling in High-Symmetry Mn12 Single-Molecule Magnets.” Polyhedron 64:128–135. doi.org/10.1016/J.Poly.2013.03.005.

Hill, S., R. S. Edwards, N. Aliaga-Alcalde, and G. Christou. 2003. “Quantum Coherence in an Exchange-Coupled Dimer of Single-Molecule Magnets.” Science 302(5647):1015–1018. doi.org/10.1126/Science.1090082.

Hill, S., R. S. Edwards, S. I. Jones, N. S. Dalal, and J. M. North. 2003. “Definitive Spectroscopic Determination of the Transverse Interactions Responsible for the Magnetic Quantum Tunneling in Mn12-Acetate.” Physical Review Letters 90(21):217204. doi.org/10.1103/Physrevlett.90.217204.

Hiscock, H. G., S. Worster, D. R. Kattnig, C. Steers, Y. Jin, D. E. Manolopoulos, H. Mouritsen, and P. J. Hore. 2016. “The Quantum Needle of the Avian Magnetic Compass.” Proceedings of the National Academy of Sciences USA 113(17): 4634–4639. doi.org/10.1073/Pnas.1600341113.

Hla, S.-W., and K.-H. Rieder. 2003. “STM Control of Chemical Reactions: Single-Molecule Synthesis.” Annual Review of Physical Chemistry 54(1):307–330. doi.org/10.1146/Annurev.Physchem.54.011002.103852.

Holland, C. M., Y. Lu, and L. W. Cheuk. 2022. “On-Demand Entanglement of Molecules in a Reconfigurable Optical Tweezer Array.” ArXiv preprint. doi.org/Arxiv:2210.06309.

Hong, C. K., Z. Y. Ou, and L. Mandel. 1987. “Measurement of Subpicosecond Time Intervals Between Two Photons by Interference.” Physical Review Letters 59(18):2044–2046. doi.org/10.1103/Physrevlett.59.2044.

Hoover, E. E., and J. A. Squier. 2013. “Advances in Multiphoton Microscopy Technology.” Nature Photonics 7(2):93–101. doi.org/10.1038/Nphoton.2012.361.

Hu, Z., B.-W. Dong, Z. Liu, J.-J. Liu, J. Su, C. Yu, J. Xiong, D.-E. Shi, Y. Wang, B.-W. Wang, A. Ardavan, Z. Shi, S.-D. Jiang, and S. Gao. 2018. “Endohedral Metallofullerene as Molecular High Spin Qubit: Diverse Rabi Cycles in Gd2@C79N.” Journal of the American Chemical Society 140(3):1123–1130. doi.org/10.1021/Jacs.7b12170.

Hussain, R., G. Allodi, A. Chiesa, E. Garlatti, D. Mitcov, A. Konstantatos, K. S. Pedersen, R. De Renzi, S. Piligkos, and S. Carretta. 2018. “Coherent Manipulation of a Molecular Ln-Based Nuclear Qudit Coupled to an Electron Qubit.” Journal of the American Chemical Society 140(31):9814–9818. doi.org/10.1021/Jacs.8b05934.

Imtiaz, A., T. M. Wallis, and P. Kabos. 2014. “Near-Field Scanning Microwave Microscopy: An Emerging Research Tool for Nanoscale Metrology.” IEEE Microwave Magazine 15(1):52–64. doi.org/10.1109/MMM.2013.2288711.

Ishizaki, A., and G. R. Fleming. 2009. “Theoretical Examination of Quantum Coherence in a Photosynthetic System at Physiological Temperature.” Proceedings of the National Academy of Sciences USA 106(41):17255–17260. doi.org/10.1073/Pnas.0908989106.

Ispasoiu, R. G., and T. Goodson. 2000. “Photon-Number Squeezing by Two-Photon Absorption in an Organic Polymer.” Optics Communications 178(4):371–376. doi.org/10.1016/S0030-4018(00)00681-7.

Ivanov, M. V., F. H. Bangerter, P. Wójcik, and A. I. Krylov. 2020. “Toward Ultracold Organic Chemistry: Prospects of Laser Cooling Large Organic Molecules.” Physical Review Letters 11(16):6670–6676. doi.org/10.1021/Acs.Jpclett.0c01960.

Jackson, C. E., C.-Y. Lin, S. H. Johnson, J. Van Tol, and J. M. Zadrozny. 2019. “Nuclear-Spin-Pattern Control of Electron-Spin Dynamics in a Series of V(IV) Complexes.” Chemical Science 10(36):8447–8454. doi.org/10.1039/C9SC02899D.

Jelezko, F., and J. Wrachtrup. 2006. “Single Defect Centres in Diamond: A Review.” Physica Status Solidi (A) 203(13): 3207–3225. doi.org/10.1002/Pssa.200671403.

Jenkins, M., T. Hümmer, M. J. Martínez-Pérez, J. García-Ripoll, D. Zueco, and F. Luis. 2013. “Coupling Single-Molecule Magnets to Quantum Circuits.” New Journal of Physics 15(9):095007. doi.org/10.1088/1367-2630/15/9/095007.

Jeschke, G. 2018. “Dipolar Spectroscopy–Double-Resonance Methods.” eMagRes 5:1459–1476.

Jo, M.-H., J. E. Grose, K. Baheti, M. M. Deshmukh, J. J. Sokol, E. M. Rumberger, D. N. Hendrickson, J. R. Long, H. Park, and D. C. Ralph. 2006. “Signatures of Molecular Magnetism in Single-Molecule Transport Spectroscopy.” Nano Letters 6(9):2014–2020. doi.org/10.1021/Nl061212i.

Johnson, S. H., C. E. Jackson, and J. M. Zadrozny. 2020. “Programmable Nuclear-Spin Dynamics in Ti(IV) Coordination Complexes.” Inorganic Chemistry 59(11):7479–7486. doi.org/10.1021/Acs.Inorgchem.0c00244.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Jones, C. G., M. W. Martynowycz, J. Hattne, T. J. Fulton, B. M. Stoltz, J. A. Rodriguez, H. M. Nelson, and T. Gonen. 2018. “The CryoEM Method MicroED as a Powerful Tool for Small Molecule Structure Determination.” ACS Central Science 4(11):1587–1592. doi.org/10.1021/Acscentsci.8b00760.

Julien, M. H., Z. H. Jang, A. Lascialfari, F. Borsa, M. Horvatić, A. Caneschi, and D. Gatteschi. 1999. “Proton NMR for Measuring Quantum Level Crossing in the Magnetic Molecular Ring Fe10.” Physical Review Letters 83(1):227–230. doi.org/10.1103/Physrevlett.83.227.

Kalb, N., A. A. Reiserer, P. C. Humphreys, J. J. W. Bakermans, S. J. Kamerling, N. H. Nickerson, S. C. Benjamin, D. J. Twitchen, M. Markham, and R. Hanson. 2017. “Entanglement Distillation Between Solid-State Quantum Network Nodes.” Science 356(6341):928–932. doi.org/10.1126/Science.Aan0070.

Kang, G., K. N. Avanaki, M. A. Mosquera, R. K. Burdick, J. P. Villabona-Monsalve, T. Goodson, III, and G. C. Schatz. 2020. “Efficient Modeling of Organic Chromophores for Entangled Two-Photon Absorption.” Journal of the American Chemical Society 142(23):10446–10458. doi.org/10.1021/Jacs.0c02808.

Kawaguchi, R., K. Hashimoto, T. Kakudate, K. Katoh, M. Yamashita, and T. Komeda. 2023. “Spatially Resolving Electron Spin Resonance of π-Radical in Single-Molecule Magnet.” Nano Letters 23(1):213–219. doi.org/10.1021/acs.nanolett.2c04049.

Kazmierczak, N. P., R. Mirzoyan, and R. G. Hadt. 2021. “The Impact of Ligand Field Symmetry on Molecular Qubit Coherence.” Journal of the American Chemical Society 143(42):17305–17315. doi.org/10.1021/Jacs.1c04605.

Kirschvink, J. L., M. M. Walker, and C. E. Diebel. 2001. “Magnetite-Based Magnetoreception.” Current Opinion in Neurobiology 11(4):462–467. doi.org/10.1016/S0959-4388(00)00235-X.

Kłos, J., and S. Kotochigova. 2020. “Prospects for Laser Cooling of Polyatomic Molecules with Increasing Complexity.” Physical Review Research 2(1):013384. doi.org/10.1103/Physrevresearch.2.013384.

Köhler, J., J. A. J. M. Disselhorst, M. C. J. M. Donckers, E. J. J. Groenen, J. Schmidt, and W. E. Moerner. 1993. “Magnetic Resonance of a Single Molecular Spin.” Nature 363(6426):242–244. doi.org/10.1038/363242a0.

Kojima J., and Q-V. Nguyen. 2004. “Entangled Biphoton Virtual-State Spectroscopy of the A2Σ+−X2Π System of OH.” Chemical Physics Letters 396(4–6):323–328. doi.org/10.1016/j.cplett.2004.08.051.

Kothe, G., M. Lukaschek, T. Yago, G. Link, K. L. Ivanov, and T.-S. Lin. 2021. “Initializing 214 Pure 14-Qubit Entangled Nuclear Spin States in a Hyperpolarized Molecular Solid.” Journal of Physical Chemistry Letters 12(14):3647–3654. doi.org/10.1021/Acs.Jpclett.1c00726.

Kovarik, S., R. Robles, R. Schlitz, T. S. Seifert, N. Lorente, P. Gambardella, and S. Stepanow. 2022. “Electron Paramagnetic Resonance of Alkali Metal Atoms and Dimers on Ultrathin MgO.” Nano Letters 22(10):4176–4181. doi.org/10.1021/Acs.Nanolett.2c00980.

Kozyryev, I., L. Baum, K. Matsuda, H. Boerge, and J. M. Doyle. 2016. “Radiation Pressure Force from Optical Cycling on a Polyatomic Molecule.” Journal of Physics B: Atomic, Molecular, and Optical Physics 49(13):134002. doi.org/10.1088/0953-4075/49/13/134002.

Kozyryev, I., L. Baum, K. Matsuda, and J. M. Doyle. 2016. “Proposal for Laser Cooling of Complex Polyatomic Molecules.” Chemistry Europe 17(22):3641–3648. doi.org/10.1002/Cphc.201601051.

Kozyryev, I., Z. Lasner, and J. M. Doyle. 2021. “Enhanced Sensitivity to Ultralight Bosonic Dark Matter in the Spectra of the Linear Radical SrOH.” Physical Review A 103(4):043313. doi.org/10.1103/Physreva.103.043313.

Kozyryev, I., T. C. Steimle, P. Yu, D.-T. Nguyen, and J. M. Doyle. 2019. “Determination of CaOH and CaOCH3 Vibrational Branching Ratios for Direct Laser Cooling and Trapping.” New Journal of Physics 21(5):052002. doi.org/10.1088/1367-2630/Ab19d7.

Kragskow, J. G. C., J. Marbey, C. D. Buch, J. Nehrkorn, M. Ozerov, S. Piligkos, S. Hill, and N. F. Chilton. 2022. “Analysis of Vibronic Coupling in a 4f Molecular Magnet with FIRMS.” Nature Communications 13(1):825. doi.org/10.1038/S41467-022-28352-2.

Krzyaniak, M. D., L. Kobr, B. K. Rugg, B. T. Phelan, E. A. Margulies, J. N. Nelson, R. M. Young, and M. R. Wasielewski. 2015. “Fast Photo-Driven Electron Spin Coherence Transfer: The Effect of Electron-Nuclear Hyperfine Coupling on Coherence Dephasing.” Journal of Materials Chemistry C 3(30):7962–7967. doi.org/10.1039/C5TC01446H.

Kubo, T., T. Goto, T. Koshiba, K. Takeda, and K. Awaga. 2002. “55Mn NMR in Mn12 Acetate: Hyperfine Interaction and Magnetic Relaxation of Cluster.” Physical Review B 65(22):224425. doi.org/10.1103/Physrevb.65.224425.

Kundu, K., J. Chen, S. Hoffman, J. Marbey, D. Komijani, Y. Duan, A. Gaita-Ariño, J. Stanton, X. Zhang, H.-P. Cheng, and S. Hill. 2023. “Electron-Nuclear Decoupling at a Spin Clock Transition.” Communications Physics 6(1):38. doi.org/10.1038/s42005-023-01152-w.

Kundu, K., J. R. K. White, S. A. Moehring, J. M. Yu, J. W. Ziller, F. Furche, W. J. Evans, and S. Hill. 2022. “A 9.2-GHz Clock Transition in a Lu(II) Molecular Spin Qubit Arising from a 3,467-MHz Hyperfine Interaction.” Nature Chemistry 14(4):392–397. doi.org/10.1038/S41557-022-00894-4.

Künstner, S., A. Chu, K. P. Dinse, A. Schnegg, J. E. Mcpeak, B. Naydenov, J. Anders, and K. Lips. 2021. “Rapid-Scan Electron Paramagnetic Resonance Using an EPR-on-a-Chip Sensor.” Magnetic Resonance 2(2):673–687. doi.org/10.5194/mr-2-673-2021.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Kutas, M., B. Haase, P. Bickert, F. Riexinger, D. Molter, and G. von Freymann. 2020. “Terahertz Quantum Sensing.” Science Advances 6(11):eaaz8065. doi.org/10.1126/sciadv.aaz8065.

Lecoq, J., N. Orlova, and B. F. Grewe. 2019. “Wide. Fast. Deep: Recent Advances in Multiphoton Microscopy of In Vivo Neuronal Activity.” Journal of Neuroscience 39(46):9042–9052. doi.org/10.1523/Jneurosci.1527-18.2019.

Lee, D.-I., and T. Goodson. 2006. “Entangled Photon Absorption in an Organic Porphyrin Dendrimer.” Journal of Physical Chemistry B 110(51):25582–25585. doi.org/10.1021/Jp066767g.

Leibfried, D. 2012. “Quantum State Preparation and Control of Single Molecular Ions.” New Journal of Physics 14(2):023029. doi.org/10.1088/1367-2630/14/2/023029.

Leisegang, M., A. Christ, S. Haldar, S. Heinze, and M. Bode. 2021. “Molecular Chains: Arranging and Programming Logic Gates.” Nano Letters 21(1):550–555. doi.org/10.1021/Acs.Nanolett.0c03984.

Lelek, M., M. T. Gyparaki, G. Beliu, F. Schueder, J. Griffié, S. Manley, R. Jungmann, M. Sauer, M. Lakadamyali, and C. Zimmer. 2021. “Single-Molecule Localization Microscopy.” Nature Reviews Methods Primers 1(1):39. doi.org/10.1038/S43586-021-00038-X.

Lemos, G. B., V. Borish, G. D. Cole, S. Ramelow, R. Lapkiewicz, and A. Zeilinger. 2014. “Quantum Imaging with Undetected Photons.” Nature 512(7515):409–412. doi.org/10.1038/nature13586.

Leuenberger, M. N., and D. Loss. 2001. “Quantum Computing in Molecular Magnets.” Nature 410(6830):789–793. doi.org/10.1038/35071024.

Li, T., F. Li, C. Altuzarra, A. Classen, and G. S. Agarwal. 2020. “Squeezed Light Induced Two-Photon Absorption Fluorescence of Fluorescein Biomarkers.” Applied Physics Letters 116(25):254001. doi.org/10.1063/5.0010909.

Li, Y., B. Shen, S. Li, Y. Zhao, J. Qu, and L. Liu. 2021. “Review of Stimulated Raman Scattering Microscopy Techniques and Applications in the Biosciences.” Advanced Biology 5(1):2000184. doi.org/10.1002/Adbi.202000184.

Lim, H. 2019. “Harmonic Generation Microscopy 2.0: New Tricks Empowering Intravital Imaging for Neuroscience.” Frontiers in Molecular Biosciences 6:99. doi.org/10.3389/Fmolb.2019.00099.

Lin, Y., D. R. Leibrandt, D. Leibfried, and C.-W. Chou. 2020. “Quantum Entanglement Between an Atom and a Molecule.” Nature 581(7808):273–277. doi.org/10.1038/S41586-020-2257-1.

Lindner, C., S. Wolf, J. Kiessling, and F. Kühnemann. 2020. “Fourier Transform Infrared Spectroscopy with Visible Light.” Optics Express 28(4):4426–4432. doi.org/10.1364/oe.382351.

Liu, J., J. Mrozek, W. K. Myers, G. A. Timco, R. E. P. Winpenny, B. Kintzel, W. Plass, and A. Ardavan. 2019. “Electric Field Control of Spins in Molecular Magnets.” Physical Review Letters 122(3):037202. doi.org/10.1103/Physrevlett.122.037202.

Liu, J., J. Mrozek, A. Ullah, Y. Duan, J. J. Baldoví, E. Coronado, A. Gaita-Ariño, and A. Ardavan. 2021. “Quantum Coherent Spin–Electric Control in a Molecular Nanomagnet at Clock Transitions.” Nature Physics 17(11):1205–1209. doi.org/10.1038/S41567-021-01355-4.

Liu, K. S., A. Henning, M. W. Heindl, R. D. Allert, J. D. Bartl, I. D. Sharp, R. Rizzato, and D. B. Bucher. 2022. “Surface NMR Using Quantum Sensors in Diamond.” Proceedings of the National Academy of Sciences USA 119(5):E2111607119. doi.org/10.1073/Pnas.2111607119.

Liu, Y., and K.-K. Ni. 2022. “Bimolecular Chemistry in the Ultracold Regime.” Annual Review of Physical Chemistry 73(1):73–96. doi.org/10.1146/Annurev-Physchem-090419-043244.

Liu, Y., M.-G. Hu, M. A. Nichols, D. Yang, D. Xie, H. Guo, and K.-K. Ni. 2021. “Precision Test of Statistical Dynamics with State-to-State Ultracold Chemistry.” Nature 593(7859):379–384. doi.org/10.1038/S41586-021-03459-6.

Lloyd, S. 1993. “A Potentially Realizable Quantum Computer.” Science 261(5128):1569–1571. doi.org/10.1126/Science.261.5128.1569.

Loh, H., K. C. Cossel, M. C. Grau, K.-K. Ni, E. R. Meyer, J. L. Bohn, J. Ye, and E. A. Cornell. 2013. “Precision Spectroscopy of Polarized Molecules in an Ion Trap.” Science 342(6163):1220–1222. doi.org/10.1126/Science.1243683.

Loss, D., and D. P. DiVincenzo. 1998. “Quantum Computation with Quantum Dots.” Physical Review A 57(1):120–126. doi.org/10.1103/Physreva.57.120.

Lotfi, H., M. A. Hassan, M. Kern, and J. Anders. 2022. “A Compact C-Band EPR-on-a-Chip Transceiver in 130-nm SiGe BiCMOS.” 17th Conference on Ph.D Research in Microelectronics and Electronics (PRIME), Villasimius (Ca), Italy, June 12–15, 2022.

Loth, S., S. Baumann, C. P. Lutz, D. M. Eigler, and A. J. Heinrich. 2012. “Bistability in Atomic-Scale Antiferromagnets.” Science 335(6065):196–199. doi.org/10.1126/Science.1214131.

Loth, S., M. Etzkorn, C. P. Lutz, D. M. Eigler, and A. J. Heinrich. 2010. “Measurement of Fast Electron Spin Relaxation Times with Atomic Resolution.” Science 329(5999):1628–1630. doi.org/10.1126/Science.1191688.

Loudon, R., and P. L. Knight. 1987. “Squeezed Light.” Journal of Modern Optics 34(6–7):709–759. doi.org/10.1080/09500348714550721.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Lovchinsky, I., J. D. Sanchez-Yamagishi, E. K. Urbach, S. Choi, S. Fang, T. I. Andersen, K. Watanabe, T. Taniguchi, A. Bylinskii, E. Kaxiras, P. Kim, H. Park, and M. D. Lukin. 2017. “Magnetic Resonance Spectroscopy of an Atomically Thin Material Using a Single-Spin Qubit.” Science 355(6324):503–507. doi.org/10.1126/Science.Aal2538.

Lovchinsky, I., A. O. Sushkov, E. Urbach, N. P. De Leon, S. Choi, K. De Greve, R. Evans, R. Gertner, E. Bersin, C. Müller, L. Mcguinness, F. Jelezko, R. L. Walsworth, H. Park, and M. D. Lukin. 2016. “Nuclear Magnetic Resonance Detection and Spectroscopy of Single Proteins Using Quantum Logic.” Science 351(6275):836–841. doi.org/10.1126/Science.Aad8022.

Luis, F., P. J. Alonso, O. Roubeau, V. Velasco, D. Zueco, D. Aguilà, J. I. Martínez, L. A. Barrios, and G. Aromí. 2020. “A Dissymmetric [Gd2] Coordination Molecular Dimer Hosting Six Addressable Spin Qubits.” Communications Chemistry 3(1):176. doi.org/10.1038/S42004-020-00422-W.

Lunghi, A., and S. Sanvito. 2020. “The Limit of Spin Lifetime in Solid-State Electronic Spins.” Physical Review Letters 11(15):6273–6278. doi.org/10.1021/Acs.Jpclett.0c01681.

Lutz, P., R. Marx, D. Dengler, A. Kromer, and J. Van Slageren. 2013. “Quantum Coherence in a Triangular Cu3 Complex.” Molecular Physics 111(18–19):2897–2902. doi.org/10.1080/00268976.2013.826421.

Macià, F., J. Lawrence, S. Hill, J. M. Hernandez, J. Tejada, P. V. Santos, C. Lampropoulos, and G. Christou. 2008. “Spin Dynamics in Single-Molecule Magnets Combining Surface Acoustic Waves and High-Frequency Electron Paramagnetic Resonance.” Physical Review B 77(2):020403. doi.org/10.1103/Physrevb.77.020403.

Maeda, K., K. B. Henbest, F. Cintolesi, I. Kuprov, C. T. Rodgers, P. A. Liddell, D. Gust, C. R. Timmel, and P. J. Hore. 2008. “Chemical Compass Model of Avian Magnetoreception.” Nature 453(7193):387–390. doi.org/10.1038/Nature06834.

Maeda, K., A. J. Robinson, K. B. Henbest, H. J. Hogben, T. Biskup, M. Ahmad, E. Schleicher, S. Weber, C. R. Timmel, and P. J. Hore. 2012. “Magnetically Sensitive Light-Induced Reactions in Cryptochrome are Consistent with Its Proposed Role as a Magnetoreceptor.” Proceedings of the National Academy of Sciences USA 109(13):4774–4779. doi.org/10.1073/Pnas.1118959109.

Mamin, H. J., M. Kim, M. H. Sherwood, C. T. Rettner, K. Ohno, D. D. Awschalom, and D. Rugar. 2013. “Nanoscale Nuclear Magnetic Resonance with a Nitrogen-Vacancy Spin Sensor.” Science 339(6119):557–560. doi.org/10.1126/Science.1231540.

Marriott, K. E. R., L. Bhaskaran, C. Wilson, M. Medarde, S. T. Ochsenbein, S. Hill, and M. Murrie. 2015. “Pushing the Limits of Magnetic Anisotropy in Trigonal Bipyramidal Ni(II).” Chemical Science 6(12):6823–6828. doi.org/10.1039/C5SC02854J.

Maurer, P. C., G. Kucsko, C. Latta, L. Jiang, N. Y. Yao, S. D. Bennett, F. Pastawski, D. Hunger, N. Chisholm, M. Markham, D. J. Twitchen, J. I. Cirac, and M. D. Lukin. 2012. “Room-Temperature Quantum Bit Memory Exceeding One Second.” Science 336(6086):1283. doi.org/10.1126/Science.1220513.

Mauser, N., and A. Hartschuh. 2014. “Tip-Enhanced Near-Field Optical Microscopy.” Chemical Society Reviews 43(4): 1248–1262. doi.org/10.1039/C3CS60258C.

McCarron D. 2018. “Laser Cooling and Trapping Molecules.” Journal of Physics B: Atomic, Molecular and Optical Physics 51(21):212001. doi.org/10.1088/1361-6455/aadfba.

McInnes, E. J. L. 2022. “Molecular Spins Clock in.” Nature Chemistry 14(4):361–362. doi.org/10.1038/S41557-022-00919-Y.

McKemmish, L. K., R. H. McKenzie, N. S. Hush, and J. R. Reimers. 2011. “Quantum Entanglement Between Electronic and Vibrational Degrees of Freedom in Molecules.” Journal of Chemical Physics 135(24):244110. doi.org/10.1063/1.3671386.

Micotti, E., Y. Furukawa, K. Kumagai, S. Carretta, A. Lascialfari, F. Borsa, G. A. Timco, and R. E. P. Winpenny. 2006. “Local Spin Moment Distribution in Antiferromagnetic Molecular Rings Probed by NMR.” Physical Review Letters 97(26):267204. doi.org/10.1103/Physrevlett.97.267204.

Mitra, D., Z. D. Lasner, G.-Z. Zhu, C. E. Dickerson, B. L. Augenbraun, A. D. Bailey, A. N. Alexandrova, W. C. Campbell, J. R. Caram, E. R. Hudson, and J. M. Doyle. 2022. “Pathway Toward Optical Cycling and Laser Cooling of Functionalized Arenes.” Physical Review Letters 13(30):7029–7035. doi.org/10.1021/Acs.Jpclett.2c01430.

Mitra, D., N. B. Vilas, C. Hallas, L. Anderegg, B. L. Augenbraun, L. Baum, C. Miller, S. Raval, and J. M. Doyle. 2020. “Direct Laser Cooling of a Symmetric Top Molecule.” Science 369(6509):1366–1369. doi.org/10.1126/Science.Abc5357.

Moffitt, J. R., Y. R. Chemla, S. B. Smith, and C. Bustamante. 2008. “Recent Advances in Optical Tweezers.” Annual Review of Biochemistry 77:205–228. doi.org/10.1146/Annurev.Biochem.77.043007.090225.

Mola, M., S. Hill, P. Goy, and M. Gross. 2000. “Instrumentation for Millimeter-Wave Magnetoelectrodynamic Investigations of Low-Dimensional Conductors and Superconductors.” Review of Scientific Instruments 71(1):186–200. doi.org/10.1063/1.1150182.

Moneron, G., and S. W. Hell. 2009. “Two-Photon Excitation STED Microscopy.” Optics Express 17(17):14567–14573. doi.org/10.1364/OE.17.014567.

Morello, A. 2008. “Quantum Nanomagnets and Nuclear Spins: An Overview.” In Quantum Magnetism, 125–138. Amsterdam: Springer Netherlands.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Morello, A., and L. J. de Jongh. 2007. “Dynamics and Thermalization of the Nuclear Spin Bath in the Single-Molecule Magnet Mn12−ac: Test for the Theory of Spin Tunneling.” Physical Review B 76(18):184425. doi.org/10.1103/Physrevb.76.184425.

Morello, A., O. N. Bakharev, H. B. Brom, R. Sessoli, and L. J. De Jongh. 2004. “Nuclear Spin Dynamics in the Quantum Regime of a Single-Molecule Magnet.” Physical Review Letters 93(19):197202. doi.org/10.1103/Physrevlett.93.197202.

Moreno-Pineda, E., and W. Wernsdorfer. 2021. “Measuring Molecular Magnets for Quantum Technologies.” Nature Reviews Physics 3(9):645–659. doi.org/10.1038/S42254-021-00340-3.

Morley, G. W., L.-C. Brunel, and J. van Tol. 2008. “A Multifrequency High-Field Pulsed Electron Paramagnetic Resonance/Electron-Nuclear Double Resonance Spectrometer.” Review of Scientific Instruments 79(6):064703. doi.org/10.1063/1.2937630.

Mukai, Y., R. Okamoto, and S. Takeuchi. 2022. “Quantum Fourier-Transform Infrared Spectroscopy in the Fingerprint Region.” Optics Express 30(13):22624–22636. doi.org/10.1364/oe.455718.

Mukamel, S., M. Freyberger, W. Schleich, M. Bellini, A. Zavatta, G. Leuchs, C. Silberhorn, R. W. Boyd, L. L. Sánchez-Soto, A. Stefanov, M. Barbieri, A. Paterova, L. Krivitsky, S. Shwartz, K. Tamasaku, K. Dorfman, F. Schlawin, V. Sandoghdar, M. Raymer, A. Marcus, O. Varnavski, T. Goodson, Z.-Y. Zhou, B.-S. Shi, S. Asban, M. Scully, G. Agarwal, T. Peng, A. V. Sokolov, Z.-D. Zhang, M. S. Zubairy, I. A. Vartanyants, E. Del Valle, and F. Laussy. 2020. “Roadmap on Quantum Light Spectroscopy.” Journal of Physics B: Atomic, Molecular, and Optical Physics 53(7):072002. doi.org/10.1088/1361-6455/Ab69a8.

Müller, C., X. Kong, J. M. Cai, K. Melentijević, A. Stacey, M. Markham, D. Twitchen, J. Isoya, S. Pezzagna, J. Meijer, J. F. Du, M. B. Plenio, B. Naydenov, L. P. Mcguinness, and F. Jelezko. 2014. “Nuclear Magnetic Resonance Spectroscopy with Single Spin Sensitivity.” Nature Communications 5(1):4703. doi.org/10.1038/Ncomms5703.

Naaman, R., and D. H. Waldeck. 2012. “Chiral-Induced Spin Selectivity Effect.” Journal of Physical Chemistry Letters 3(16):2178–2187. doi.org/10.1021/Jz300793y.

Nagarajan, K., A. Thomas, and T. W. Ebbesen. 2021. “Chemistry under Vibrational Strong Coupling.” Journal of the American Chemical Society 143(41):16877–16889. doi.org/10.1021/jacs.1c07420.

Nehrkorn, J., S. M. Greer, B. J. Malbrecht, K. J. Anderton, A. Aliabadi, J. Krzystek, A. Schnegg, K. Holldack, C. Herrmann, T. A. Betley, S. Stoll, and S. Hill. 2021. “Spectroscopic Investigation of a Metal–Metal-Bonded Fe6 Single-Molecule Magnet with an Isolated S = 19/2 Giant-Spin Ground State.” Inorganic Chemistry 60(7):4610–4622. doi.org/10.1021/Acs.Inorgchem.0c03595.

Nelson, J. N., M. D. Krzyaniak, N. E. Horwitz, B. K. Rugg, B. T. Phelan, and M. R. Wasielewski. 2017. “Zero Quantum Coherence in a Series of Covalent Spin-Correlated Radical Pairs.” Journal of Physical Chemistry A 121(11):2241–2252. doi.org/10.1021/Acs.Jpca.7b00587.

Neugebauer, P., D. Bloos, R. Marx, P. Lutz, M. Kern, D. Aguilà, J. Vaverka, O. Laguta, C. Dietrich, R. Clérac, and J. Van Slageren. 2018. “Ultra-Broadband EPR Spectroscopy in Field and Frequency Domains.” Physical Chemistry Chemical Physics 20(22):15528–15534. doi.org/10.1039/C7CP07443C.

Nguyen, T. N., M. Shiddiq, T. Ghosh, K. A. Abboud, S. Hill, and G. Christou. 2015. “Covalently Linked Dimer of Mn3 Single-Molecule Magnets and Retention of Its Structure and Quantum Properties in Solution.” Journal of the American Chemical Society 137(22):7160–7168. doi.org/10.1021/Jacs.5b02677.

Ni, K.-K., S. Ospelkaus, M. H. G. De Miranda, A. Pe’er, B. Neyenhuis, J. J. Zirbel, S. Kotochigova, P. S. Julienne, D. S. Jin, and J. Ye. 2008. “A High Phase-Space-Density Gas of Polar Molecules.” Science 322(5899):231–235. doi.org/10.1126/Science.1163861.

Ni, K. K., S. Ospelkaus, D. J. Nesbitt, J. Ye, and D. S. Jin. 2009. “A Dipolar Gas of Ultracold Molecules.” Physical Chemistry Chemical Physics 11(42):9626–9639. doi.org/10.1039/B911779B.

NIST (National Institute of Standards and Technology). 2009. “Physicists Find Way to Control Individual Bits in Quantum Computers.” NIST, July. https://www.nist.gov/news-events/news/2009/07/physicists-find-way-control-individual-bitsquantum-computers.

Olshansky, J. H., M. D. Krzyaniak, R. M. Young, and M. R. Wasielewski. 2019. “Photogenerated Spin-Entangled Qubit (Radical) Pairs in DNA Hairpins: Observation of Spin Delocalization and Coherence.” Journal of the American Chemical Society 141(5):2152–2160. doi.org/10.1021/Jacs.8b13155.

Ospelkaus, S., K.-K. Ni, D. Wang, M. H. G. De Miranda, B. Neyenhuis, G. Quéméner, P. S. Julienne, J. L. Bohn, D. S. Jin, and J. Ye. 2010. “Quantum-State Controlled Chemical Reactions of Ultracold Potassium-Rubidium Molecules.” Science 327(5967):853–857. doi.org/10.1126/Science.1184121.

Parzuchowski, K. M., A. Mikhaylov, M. D. Mazurek, R. N. Wilson, D. J. Lum, T. Gerrits, C. H. Camp, M. J. Stevens, and R. Jimenez. 2021. “Setting Bounds on Entangled Two-Photon Absorption Cross Sections in Common Fluorophores.” Physical Review Applied 15(4):044012. doi.org/10.1103/Physrevapplied.15.044012.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Perrin, M. L., E. Burzurí, and H. S. J. van der Zant. 2015. “Single-Molecule Transistors.” Chemical Society Reviews 44(4): 902–919. doi.org/10.1039/C4CS00231H.

Pietsch, T., S. Egle, M. Keller, H. Fridtjof-Pernau, F. Strigl, and E. Scheer. 2016. “Microwave-Induced Direct Spin-Flip Transitions in Mesoscopic Pd/Co Heterojunctions.” New Journal of Physics 18(9):093045. doi.org/10.1088/1367-2630/18/9/093045.

Pinto, D., D. Paone, B. Kern, T. Dierker, R. Wieczorek, A. Singha, D. Dasari, A. Finkler, W. Harneit, J. Wrachtrup, and K. Kern. 2020. “Readout and Control of an Endofullerene Electronic Spin.” Nature Communications 11(1):6405. doi.org/10.1038/S41467-020-20202-3.

Pirandola, S., J. Eisert, C. Weedbrook, A. Furusawa, and S. L. Braunstein. 2015. “Advances in Quantum Teleportation.” Nature Photonics 9:641–652. doi.org/10.1038/Nphoton.2015.154.

Plasser, F. 2016. “Entanglement Entropy of Electronic Excitations.” Journal of Chemical Physics 144:194107. doi.org/10.1063/1.4949535.

Potočnik, A., A. Bargerbos, F. A. Y. N. Schröder, S. A. Khan, M. C. Collodo, S. Gasparinetti, Y. Salathé, C. Creatore, C. Eichler, H. E. Türeci, A. W. Chin, and A. Wallraff. 2018. “Studying Light-Harvesting Models with Superconducting Circuits.” Nature Communications 9(1):904. doi.org/10.1038/s41467-018-03312-x.

Prescimone, A., C. Morien, D. Allan, J. A. Schlueter, S. W. Tozer, J. L. Manson, S. Parsons, E. K. Brechin, and S. Hill. 2012. “Pressure-Driven Orbital Reorientations and Coordination-Sphere Reconstructions in [CuF2(H2O)2(pyz)].” Angewandte Chemie International Edition 51(30):7490–7494. doi.org/10.1002/Anie.201202367.

Price, H. 2022. “Simulating Four-Dimensional Physics in the Laboratory.” Physics Today 75(4):38–44. doi.org/10.1063/PT.3.4981.

Purcell, E. M., H. C. Torrey, and R. V. Pound. 1946. “Resonance Absorption by Nuclear Magnetic Moments in a Solid.” Physical Review 69(1–2):37–38. doi.org/10.1103/Physrev.69.37.

Qian, C., K. Miao, L.-E. Lin, X. Chen, J. Du, and L. Wei. 2021. “Super-Resolution Label-Free Volumetric Vibrational Imaging.” Nature Communications 12(1):3648. doi.org/10.1038/S41467-021-23951-X.

Ray, K., S. P. Ananthavel, D. H. Waldeck, and R. Naaman. 1999. “Asymmetric Scattering of Polarized Electrons by Organized Organic Films of Chiral Molecules.” Science 283(5403):814–816. doi.org/10.1126/Science.283.5403.814.

Raymer, M. G., A. H. Marcus, J. R. Widom, and D. L. P. Vitullo. 2013. “Entangled Photon-Pair Two-Dimensional Fluorescence Spectroscopy (EPP-2DFS).” Journal of Physical Chemistry B 117(49):15559–15575. doi.org/10.1021/Jp405829n.

Rehberg, M., F. Krombach, U. Pohl, and S. Dietzel. 2011. “Label-Free 3D Visualization of Cellular and Tissue Structures in Intact Muscle with Second and Third Harmonic Generation Microscopy.” PLoS ONE 6(11):E28237. doi.org/10.1371/Journal.Pone.0028237.

Ribeiro, R. F., L. A. Martínez-Martínez, M. Du, J. Campos-Gonzalez-Angulo, and J. Yuen-Zhou. 2018. “Polariton Chemistry: Controlling Molecular Dynamics with Optical Cavities.” Chemical Science 9(30):6325–6339. doi.org/10.1039/C8SC01043A.

Riedel, D., I. Söllner, B. J. Shields, S. Starosielec, P. Appel, E. Neu, P. Maletinsky, and R. J. Warburton. 2017. “Deterministic Enhancement of Coherent Photon Generation from a Nitrogen-Vacancy Center in Ultrapure Diamond.” Physical Review X 7(3):031040. doi.org/10.1103/Physrevx.7.031040.

Rieke, F., and D. A. Baylor. 1998. “Single-Photon Detection by Rod Cells of the Retina.” Reviews of Modern Physics 70(3):1027–1036. doi.org/10.1103/Revmodphys.70.1027.

Rismani Yazdi, S., R. Nosrati, C. A. Stevens, D. Vogel, P. L. Davies, and C. Escobedo. 2018. “Magnetotaxis Enables Magneto-tactic Bacteria to Navigate in Flow.” Small 14(5):1702982. doi.org/10.1002/Smll.201702982.

Rodgers, L. V. H., L. B. Hughes, M. Xie, P. C. Maurer, S. Kolkowitz, A. C. B. Jayich, and N. P. De Leon. 2021. “Materials Challenges for Quantum Technologies Based on Color Centers in Diamond.” MRS Bulletin 46(7):623–633. doi.org/10.1557/S43577-021-00137-W.

Röhlsberger, R., J. Evers, and S. Shwartz. 2014. “Quantum and Nonlinear Optics with Hard X-Rays.” In Synchrotron Light Sources and Free-Electron Lasers: Accelerator Physics, Instrumentation and Science Applications, edited by E. Jaeschke, S. Khan, J. R. Schneider, and J. B. Hastings, 1–28. Cham: Springer International Publishing.

Roslyak, O., C. A. Marx, and S. Mukamel. 2009. “Nonlinear Spectroscopy with Entangled Photons: Manipulating Quantum Pathways of Matter.” Physical Review A 79(3):033832. doi.org/10.1103/Physreva.79.033832.

Rosner, B. T., and D. W. van der Weide. 2002. “High-Frequency Near-Field Microscopy.” Review of Scientific Instruments 73(7):2505–2525. doi.org/10.1063/1.1482150.

Rossini, A. J., A. Zagdoun, M. Lelli, A. Lesage, C. Copéret, and L. Emsley. 2013. “Dynamic Nuclear Polarization Surface Enhanced NMR Spectroscopy.” Accounts of Chemical Research 46(9):1942–1951. doi.org/10.1021/Ar300322x.

Ruamps, R., R. Maurice, L. Batchelor, M. Boggio-Pasqua, R. Guillot, A. L. Barra, J. Liu, E.-E. Bendeif, S. Pillet, S. Hill, T. Mallah, and N. Guihéry. 2013. “Giant Ising-Type Magnetic Anisotropy in Trigonal Bipyramidal Ni(II) Complexes: Experiment and Theory.” Journal of the American Chemical Society 135(8):3017–3026. doi.org/10.1021/Ja308146e.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Rugar, D., R. Budakian, H. J. Mamin, and B. W. Chui. 2004. “Single Spin Detection by Magnetic Resonance Force Microscopy.” Nature 430(6997):329–332. doi.org/10.1038/Nature02658.

Rugar, D., H. J. Mamin, M. H. Sherwood, M. Kim, C. T. Rettner, K. Ohno, and D. D. Awschalom. 2015. “Proton Magnetic Resonance Imaging Using a Nitrogen–Vacancy Spin Sensor.” Nature Nanotechnology 10(2):120–124. doi.org/10.1038/Nnano.2014.288.

Rugg, B. K., M. D. Krzyaniak, B. T. Phelan, M. A. Ratner, R. M. Young, and M. R. Wasielewski. 2019. “Photodriven Quantum Teleportation of an Electron Spin State in a Covalent Donor-Acceptor-Radical System.” Nature Chemistry 11(11):981–986. doi.org/10.1038/S41557-019-0332-8.

Saeedi, K., S. Simmons, J. Z. Salvail, P. Dluhy, H. Riemann, N. V. Abrosimov, P. Becker, H.-J. Pohl, J. J. L. Morton, and M. L. W. Thewalt. 2013. “Room-Temperature Quantum Bit Storage Exceeding 39 Minutes Using Ionized Donors in Silicon-28.” Science 342(6160):830–833. doi.org/10.1126/Science.1239584.

Saleh, B. E. A., B. M. Jost, H.-B. Fei, and M. C. Teich. 1998. “Entangled-Photon Virtual-State Spectroscopy.” Physical Review Letters 80(16):3483–3486. doi.org/10.1103/Physrevlett.80.3483.

Salman, Z., K. H. Chow, R. I. Miller, A. Morello, T. J. Parolin, M. D. Hossain, T. A. Keeler, C. D. P. Levy, W. A. Macfarlane, G. D. Morris, H. Saadaoui, D. Wang, R. Sessoli, G. G. Condorelli, and R. F. Kiefl. 2007. “Local Magnetic Properties of a Monolayer of Mn12 Single Molecule Magnets.” Nano Letters 7(6):1551–1555. doi.org/10.1021/Nl070366a.

Sarovar, M., A. Ishizaki, G. R. Fleming, and K. B. Whaley. 2010. “Quantum Entanglement in Photosynthetic Light-Harvesting Complexes.” Nature Physics 6(6):462–467. doi.org/10.1038/Nphys1652.

Sato, K., S. Nakazawa, R. Rahimi, T. Ise, S. Nishida, T. Yoshino, N. Mori, K. Toyota, D. Shiomi, Y. Yakiyama, Y. Morita, M. Kitagawa, K. Nakasuji, M. Nakahara, H. Hara, P. Carl, P. Höfer, and T. Takui. 2009. “Molecular Electron-Spin Quantum Computers and Quantum Information Processing: Pulse-Based Electron Magnetic Resonance Spin Technology Applied to Matter Spin-Qubits.” Journal of Materials Chemistry 19(22):3739–3754. doi.org/10.1039/B819556K.

Sato, K., R. Rahimi, N. Mori, S. Nishida, K. Toyota, D. Shiomi, Y. Morita, A. Ueda, S. Suzuki, K. Furukawa, T. Nakamura, M. Kitagawa, K. Nakasuji, M. Nakahara, H. Hara, P. Carl, P. Höfer, and T. Takui. 2007. “Implementation of Molecular Spin Quantum Computing by Pulsed ENDOR Technique: Direct Observation of Quantum Entanglement and Spinor.” Physica E: Low-Dimensional Systems and Nanostructures 40(2):363–366. doi.org/10.1016/J.Physe.2007.06.031.

Schlawin, F., K. E. Dorfman, B. P. Fingerhut, and S. Mukamel. 2013. “Suppression of Population Transport and Control of Exciton Distributions by Entangled Photons.” Nature Communications 4(1):1782. doi.org/10.1038/Ncomms2802.

Schlawin, F., K. E. Dorfman, and S. Mukamel. 2018. “Entangled Two-Photon Absorption Spectroscopy.” Accounts of Chemical Research 51(9):2207–2214. doi.org/10.1021/Acs.Accounts.8b00173.

Schlegel, C., J. Van Slageren, M. Manoli, E. K. Brechin, and M. Dressel. 2008. “Direct Observation of Quantum Coherence in Single-Molecule Magnets.” Physical Review Letters 101(14):147203. doi.org/10.1103/Physrevlett.101.147203.

Schnegg, A., J. Behrends, K. Lips, R. Bittl, and K. Holldack. 2009. “Frequency Domain Fourier Transform THz-EPR on Single Molecule Magnets Using Coherent Synchrotron Radiation.” Physical Chemistry Chemical Physics 11(31):6820–6825. doi.org/10.1039/B905745E.

Scholes, G. D., G. R. Fleming, L. X. Chen, A. Aspuru-Guzik, A. Buchleitner, D. F. Coker, G. S. Engel, R. Van Grondelle, A. Ishizaki, D. M. Jonas, J. S. Lundeen, J. K. Mccusker, S. Mukamel, J. P. Ogilvie, A. Olaya-Castro, M. A. Ratner, F. C. Spano, K. B. Whaley, and X. Zhu. 2017. “Using Coherence to Enhance Function in Chemical and Biophysical Systems.” Nature 543(7647):647–656. doi.org/10.1038/Nature21425.

Scholl, P., M. Schuler, H. J. Williams, A. A. Eberharter, D. Barredo, K.-N. Schymik, V. Lienhard, L.-P. Henry, T. C. Lang, T. Lahaye, A. M. Läuchli, and A. Browaeys. 2021. “Quantum Simulation of 2D Antiferromagnets with Hundreds of Rydberg Atoms.” Nature 595:233–238. doi.org/10.1038/s41586-021-03585-1.

Scholten, S. C., A. J. Healey, I. O. Robertson, G. J. Abrahams, D. A. Broadway, and J.-P. Tetienne. 2021. “Widefield Quantum Microscopy with Nitrogen-Vacancy Centers in Diamond: Strengths, Limitations, and Prospects.” Journal of Applied Physics 130(15):150902. doi.org/10.1063/5.0066733.

Schrödinger, E. 2012. What Is Life? With Mind and Matter and Autobiographical Sketches, Canto Classics. Cambridge: Cambridge University Press.

Seo, C., and T.-T. Kim. 2022. “Terahertz Near-Field Spectroscopy for Various Applications.” Journal of the Korean Physical Society 81:549–561. doi.org/10.1007/S40042-022-00404-2.

Sessoli, R., D. Gatteschi, A. Caneschi, and M. A. Novak. 1993. “Magnetic Bistability in a Metal-Ion Cluster.” Nature 365(6442):141–143. doi.org/10.1038/365141a0.

Sessoli, R., H. L. Tsai, A. R. Schake, S. Wang, J. B. Vincent, K. Folting, D. Gatteschi, G. Christou, and D. N. Hendrickson. 1993. “High-Spin Molecules: [Mn12O12(O2CR)16(H2O)4].” Journal of the American Chemical Society 115(5):1804–1816. doi.org/10.1021/Ja00058a027.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Shapiro, D. A., S. Babin, R. S. Celestre, W. Chao, R. P. Conley, P. Denes, B. Enders, P. Enfedaque, S. James, J. M. Joseph, H. Krishnan, S. Marchesini, K. Muriki, K. Nowrouzi, S. R. Oh, H. Padmore, T. Warwick, L. Yang, V. V. Yashchuk, Y.-S. Yu, and J. Zhao. 2020. “An Ultrahigh-Resolution Soft X-Ray Microscope for Quantitative Analysis of Chemically Heterogeneous Nanomaterials.” Science Advances 6(51):Eabc4904. doi.org/doi:10.1126/Sciadv.Abc4904.

Shashkova, S., and M. C. Leake. 2017. “Single-Molecule Fluorescence Microscopy Review: Shedding New Light on Old Problems.” Bioscience Reports 37(4). doi.org/10.1042/Bsr20170031.

Shi, L., A. A. Fung, and A. Zhou. 2021. “Advances in Stimulated Raman Scattering Imaging for Tissues and Animals.” Quantitative Imaging in Medicine and Surgery 11(3):1078–1101. doi.org/10.21037/Qims-20-712.

Shi, M., P. F. Herskind, M. Drewsen, and I. L. Chuang. 2013. “Microwave Quantum Logic Spectroscopy and Control of Molecular Ions.” New Journal of Physics 15(11):113019. doi.org/10.1088/1367-2630/15/11/113019.

Shiddiq, M., D. Komijani, Y. Duan, A. Gaita-Ariño, E. Coronado, and S. Hill. 2016. “Enhancing Coherence in Molecular Spin Qubits Via Atomic Clock Transitions.” Nature 531(7594):348–351. doi.org/10.1038/Nature16984.

Shuman, E. S., J. F. Barry, D. R. Glenn, and D. Demille. 2009. “Radiative Force from Optical Cycling on a Diatomic Molecule.” Physical Review Letters 103(22):223001. doi.org/10.1103/Physrevlett.103.223001.

Shwartz, S., R. N. Coffee, J. M. Feldkamp, Y. Feng, J. B. Hastings, G. Y. Yin, and S. E. Harris. 2012. “X-Ray Parametric Down-Conversion in the Langevin Regime.” Physical Review Letters 109(1):013602. doi.org/10.1103/Physrevlett.109.013602.

Simmons, S., R. M. Brown, H. Riemann, N. V. Abrosimov, P. Becker, H.-J. Pohl, M. L. W. Thewalt, K. M. Itoh, and J. J. L. Morton. 2011. “Entanglement in a Solid-State Spin Ensemble.” Nature 470(7332):69–72. doi.org/10.1038/Nature09696.

Smith, T. A., Z. Wang, and Y. Shih. 2020. “Two-Photon X-Ray Ghost Microscope.” Optics Express 28(22):32249–32265. doi.org/10.1364/OE.401449.

Sofer, S., O. Sefi, E. Strizhevsky, H. Aknin, S. P. Collins, G. Nisbet, B. Detlefs, Ch. J. Sahle, and S. Shwartz. 2019. “Observation of Strong Nonlinear Interactions in Parametric Down-Conversion of X-Rays into Ultraviolet Radiation.” Nature Communications 10(1):5673. doi.org/10.1038/S41467-019-13629-W.

Squire, R. H., N. H. March, R. A. Minnick, and R. Turschmann. 2013. “Comparison of Various Types of Coherence and Emergent Coherent Systems.” International Journal of Quantum Chemistry 113(19):2181–2199. doi.org/10.1002/Qua.24423.

Stamp, P. C. E., and I. S. Tupitsyn. 2004. “Coherence Window in the Dynamics of Quantum Nanomagnets.” Physical Review B 69(1):014401. doi.org/10.1103/Physrevb.69.014401.

Stasiw, D. E., J. Zhang, G. Wang, R. Dangi, B. W. Stein, D. A. Shultz, M. L. Kirk, L. Wojtas, and R. D. Sommer. 2015. “Determining the Conformational Landscape of ς and π Coupling Using Para-Phenylene and ‘Aviram–Ratner’ Bridges.” Journal of the American Chemical Society 137(29):9222–9225. doi.org/10.1021/Jacs.5b04629.

Staudacher, T., F. Shi, S. Pezzagna, J. Meijer, J. Du, C. A. Meriles, F. Reinhard, and J. Wrachtrup. 2013. “Nuclear Magnetic Resonance Spectroscopy on a (5-Nanometer)3 Sample Volume.” Science 339(6119):561–563. doi.org/10.1126/Science.1231675.

Stevens, M. A., J. E. Mckay, J. L. S. Robinson, H. El Mkami, G. M. Smith, and D. G. Norman. 2016. “The Use of the Rx Spin Label in Orientation Measurement on Proteins, by EPR.” Physical Chemistry Chemical Physics 18(8):5799–5806. doi.org/10.1039/C5CP04753F.

Stuhl, B. K., B. C. Sawyer, D. Wang, and J. Ye. 2008. “Magneto-Optical Trap for Polar Molecules.” Physical Review Letters 101(24):243002. doi.org/10.1103/Physrevlett.101.243002.

Subramanya, M. V. H., J. Marbey, K. Kundu, J. E. Mckay, and S. Hill. 2022. “Broadband Fourier-Transform-Detected EPR at W-Band.” Applied Magnetic Resonance 54:165–181. doi.org/10.1007/S00723-022-01499-3.

Sun, K., M. F. Gelin, and Y. Zhao. 2022. “Engineering Cavity Singlet Fission in Rubrene.” Physical Review Letters 13(18): 4090–4097. doi.org/10.1021/Acs.Jpclett.2c00801.

Sushkov, A. O., I. Lovchinsky, N. Chisholm, R. L. Walsworth, H. Park, and M. D. Lukin. 2014. “Magnetic Resonance Detection of Individual Proton Spins Using Quantum Reporters.” Physical Review Letters 113(19):197601. doi.org/10.1103/Physrevlett.113.197601.

Suturina, E. A., J. Nehrkorn, J. M. Zadrozny, J. Liu, M. Atanasov, T. Weyhermüller, D. Maganas, S. Hill, A. Schnegg, E. Bill, J. R. Long, and F. Neese. 2017. “Magneto-Structural Correlations in Pseudotetrahedral Forms of the [Co(SPh)4]2– Complex Probed by Magnetometry, MCD Spectroscopy, Advanced EPR Techniques, and Ab Initio Electronic Structure Calculations.” Inorganic Chemistry 56(5):3102–3118. doi.org/10.1021/Acs.Inorgchem.7b00097.

Svidzinsky, A., G. Agarwal, A. Classen, A. V. Sokolov, A. Zheltikov, M. S. Zubairy, and M. O. Scully. 2021. “Enhancing Stimulated Raman Excitation and Two-Photon Absorption Using Entangled States of Light.” Physical Review Research 3(4):043029. doi.org/10.1103/Physrevresearch.3.043029.

Szoke, S., M. He, B. P. Hickam, and S. K. Cushing. 2021. “Designing High-Power, Octave Spanning Entangled Photon Sources for Quantum Spectroscopy.” Journal of Chemical Physics 154(24):244201. doi.org/10.1063/5.0053688.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Szoke, S., H. Liu, B. P. Hickam, M. He, and S. K. Cushing. 2020. “Entangled Light–Matter Interactions and Spectroscopy.” Journal of Materials Chemistry C 8(31):10732–10741. doi.org/10.1039/D0tc02300k.

Tabakaev, D., A. Djorović, L. La Volpe, G. Gaulier, S. Ghosh, L. Bonacina, J. P. Wolf, H. Zbinden, and R. T. Thew. 2022. “Spatial Properties of Entangled Two-Photon Absorption.” Physical Review Letters 129(18):183601. doi.org/10.1103/Physrevlett.129.183601.

Tabakaev, D., M. Montagnese, G. Haack, L. Bonacina, J. P. Wolf, H. Zbinden, and R. T. Thew. 2021. “Energy-Time-Entangled Two-Photon Molecular Absorption.” Physical Review A 103(3):033701. doi.org/10.1103/Physreva.103.033701.

Takahashi, S., and S. Hill. 2005. “Rotating Cavity for High-Field Angle-Dependent Microwave Spectroscopy of Low-Dimensional Conductors and Magnets.” Review of Scientific Instruments 76(2):023114. doi.org/10.1063/1.1852859.

Takahashi, S., R. Hanson, J. Van Tol, M. S. Sherwin, and D. D. Awschalom. 2008. “Quenching Spin Decoherence in Diamond Through Spin Bath Polarization.” Physical Review Letters 101(4):047601. doi.org/10.1103/Physrevlett.101.047601.

Takahashi, S., I. S. Tupitsyn, J. Van Tol, C. C. Beedle, D. N. Hendrickson, and P. C. E. Stamp. 2011. “Decoherence in Crystals of Quantum Molecular Magnets.” Nature 476(7358):76–79. doi.org/10.1038/Nature10314.

Takahashi, S., J. Van Tol, C. C. Beedle, D. N. Hendrickson, L.-C. Brunel, and M. S. Sherwin. 2009. “Coherent Manipulation and Decoherence of S=10 Single-Molecule Magnets.” Physical Review Letters 102(8):087603. doi.org/10.1103/Physrevlett.102.087603.

Tegmark, M. 2000. “Importance of Quantum Decoherence in Brain Processes.” Physical Review E 61(4):4194–4206. doi.org/10.1103/Physreve.61.4194.

Thiele, S., F. Balestro, R. Ballou, S. Klyatskaya, M. Ruben, and W. Wernsdorfer. 2014. “Electrically Driven Nuclear Spin Resonance in Single-Molecule Magnets.” Science 344(6188):1135–1138. doi.org/10.1126/Science.1249802.

Thirunavukkuarasu, K., S. M. Winter, C. C. Beedle, A. E. Kovalev, R. T. Oakley, and S. Hill. 2015. “Pressure Dependence of the Exchange Anisotropy in an Organic Ferromagnet.” Physical Review B 91(1):014412. doi.org/10.1103/Physrevb.91.014412.

Thomann, H., T. V. Morgan, H. Jin, S. J. N. Burgmayer, R. E. Bare, and E. I. Stiefel. 1987. “Protein Nitrogen Coordination to the Iron-Molybdenum Center of Nitrogenase from Clostridium Pasteurianum.” Journal of the American Chemical Society 109(25):7913–7914. doi.org/10.1021/Ja00259a067.

Togan, E., Y. Chu, A. S. Trifonov, L. Jiang, J. Maze, L. Childress, M. V. G. Dutt, A. S. Sørensen, P. R. Hemmer, A. S. Zibrov, and M. D. Lukin. 2010. “Quantum Entanglement Between an Optical Photon and a Solid-State Spin Qubit.” Nature 466(7307):730–734. doi.org/10.1038/Nature09256.

Truppe, S., H. J. Williams, M. Hambach, L. Caldwell, N. J. Fitch, E. A. Hinds, B. E. Sauer, and M. R. Tarbutt. 2017. “Molecules Cooled Below the Doppler Limit.” Nature Physics 13(12):1173–1176. doi.org/10.1038/Nphys4241.

Turin, L. 1996. “A Spectroscopic Mechanism for Primary Olfactory Reception.” Chemical Senses 21(6):773–791. doi.org/10.1093/Chemse/21.6.773.

Uber, J. S., M. Estrader, J. Garcia, P. Lloyd-Williams, A. Sadurní, D. Dengler, J. Van Slageren, N. F. Chilton, O. Roubeau, S. J. Teat, J. Ribas-Ariño, and G. Aromí. 2017. “Molecules Designed to Contain Two Weakly Coupled Spins with a Photoswitchable Spacer.” Chemistry: A European Journal 23(55):13648–13659. doi.org/10.1002/Chem.201702171.

Ullah, A., Z. Hu, J. Cerdá, J. Aragó, and A. Gaita-Ariño. 2022. “Electrical Two-Qubit Gates within a Pair of Clock-Qubit Magnetic Molecules.” npj Quantum Information 8(1). https://doi.org/10.1038/s41534-022-00647-8.

Upton, L., M. Harpham, O. Suzer, M. Richter, S. Mukamel, and T. Goodson, III. 2013. “Optically Excited Entangled States in Organic Molecules Illuminate the Dark.” Physical Review Letters 4(12):2046–2052. doi.org/10.1021/Jz400851d.

Van Doorslaer, S. 2017. “Hyperfine Spectroscopy: ESEEM.” eMagRes 6:51–70.

van Slageren, J. 2019. “Spin–Electric Coupling.” Nature Materials 18(4):300–301. doi.org/10.1038/S41563-019-0314-7.

van Slageren, J., S. Vongtragool, B. Gorshunov, A. A. Mukhin, N. Karl, J. Krzystek, J. Telser, A. Müller, C. Sangregorio, D. Gatteschi, and M. Dressel. 2003. “Frequency-Domain Magnetic Resonance Spectroscopy of Molecular Magnetic Materials.” Physical Chemistry Chemical Physics 5(18):3837–3843. doi.org/10.1039/B305328H.

van Tol, J., L.-C. Brunel, and R. J. Wylde. 2005. “A Quasioptical Transient Electron Spin Resonance Spectrometer Operating at 120 and 240 GHz.” Review of Scientific Instruments 76(7):074101. doi.org/10.1063/1.1942533.

Varnavski, O., and T. Goodson, III. 2020. “Two-Photon Fluorescence Microscopy at Extremely Low Excitation Intensity: The Power of Quantum Correlations.” Journal of the American Chemical Society 142(30):12966–12975. doi.org/10.1021/Jacs.0c01153.

Varnavski, O., C. Gunthardt, A. Rehman, G. D. Luker, and T. Goodson, III. 2022. “Quantum Light-Enhanced Two-Photon Imaging of Breast Cancer Cells.” Physical Review Letters 13(12):2772–2781. doi.org/10.1021/Acs.Jpclett.2c00695.

Varnavski, O., B. Pinsky, and T. Goodson, III. 2017. “Entangled Photon Excited Fluorescence in Organic Materials: An Ultrafast Coincidence Detector.” Physical Review Letters 8(2):388–393. doi.org/10.1021/Acs.Jpclett.6b02378.

Vilas, N. B., C. Hallas, L. Anderegg, P. Robichaud, A. Winnicki, D. Mitra, and J. M. Doyle. 2022. “Magneto-Optical Trapping and Sub-Doppler Cooling of a Polyatomic Molecule.” Nature 606(7912):70–74. doi.org/10.1038/S41586-022-04620-5.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Villabona-Monsalve, J. P., O. Varnavski, B. A. Palfey, and T. Goodson, III. 2018. “Two-Photon Excitation of Flavins and Flavoproteins with Classical and Quantum Light.” Journal of the American Chemical Society 140(44):14562–14566. doi.org/10.1021/Jacs.8b08515.

Vincent, R., S. Klyatskaya, M. Ruben, W. Wernsdorfer, and F. Balestro. 2012. “Electronic Read-Out of a Single Nuclear Spin Using a Molecular Spin Transistor.” Nature 488(7411):357–360. doi.org/10.1038/Nature11341.

Volkovich, S., and S. Shwartz. 2020. “Subattosecond X-Ray Hong-Ou-Mandel Metrology.” Optics Letters 45(10):2728–2731. doi.org/10.1364/Ol.382044.

Wang, H., and W. Xiong. 2021. “Vibrational Sum-Frequency Generation Hyperspectral Microscopy for Molecular Self-Assembled Systems.” Annual Review of Physical Chemistry 72(1):279–306. doi.org/10.1146/Annurev-Physchem-090519-050510.

Wang, X., J. E. Mckay, B. Lama, J. Van Tol, T. Li, K. Kirkpatrick, Z. Gan, S. Hill, J. R. Long, and H. C. Dorn. 2018. “Gadolinium Based Endohedral Metallofullerene Gd2@C79N as a Relaxation Boosting Agent for Dissolution DNP at High Fields.” Chemical Communications 54(19):2425–2428. doi.org/10.1039/C7CC09765D.

Wang, Z., S. Datta, C. Papatriantafyllopoulou, G. Christou, N. S. Dalal, J. Van Tol, and S. Hill. 2011. “Spin Decoherence in an Iron-Based Magnetic Cluster.” Polyhedron 30(18):3193–3196. doi.org/10.1016/J.Poly.2011.04.009.

Warner, M., S. Din, I. S. Tupitsyn, G. W. Morley, A. M. Stoneham, J. A. Gardener, Z. Wu, A. J. Fisher, S. Heutz, C. W. M. Kay, and G. Aeppli. 2013. “Potential for Spin-Based Information Processing in a Thin-Film Molecular Semiconductor.” Nature 503(7477):504–508. doi.org/10.1038/Nature12597.

Wasielewski, M. R., M. D. E. Forbes, N. L. Frank, K. Kowalski, G. D. Scholes, J. Yuen-Zhou, M. A. Baldo, D. E. Freedman, R. H. Goldsmith, T. Goodson, M. L. Kirk, J. K. Mccusker, J. P. Ogilvie, D. A. Shultz, S. Stoll, and K. Birgitta Whaley. 2020. “Exploiting Chemistry and Molecular Systems for Quantum Information Science.” Nature Reviews Chemistry 4(9):490–504. doi.org/10.1038/S41570-020-0200-5.

Weiden, N., H. Käss, and K. P. Dinse. 1999. “Pulse Electron Paramagnetic Resonance (EPR) and Electron−Nuclear Double Resonance (ENDOR) Investigation of N@C60 in Polycrystalline C60.” Journal of Physical Chemistry B 103(45): 9826–9830. doi.org/10.1021/Jp9914471.

Weijers, H. W., W. D. Markiewicz, A. V. Gavrilin, A. J. Voran, Y. L. Viouchkov, S. R. Gundlach, P. D. Noyes, D. V. Abraimov, H. Bai, S. T. Hannahs, and T. P. Murphy. 2016. “Progress in the Development and Construction of a 32-T Superconducting Magnet.” IEEE Transactions on Applied Superconductivity 26(4):1–7. doi.org/10.1109/TASC.2016.2517022.

Williams, H. J., L. Caldwell, N. J. Fitch, S. Truppe, J. Rodewald, E. A. Hinds, B. E. Sauer, and M. R. Tarbutt. 2018. “Magnetic Trapping and Coherent Control of Laser-Cooled Molecules.” Physical Review Letters 120(6):163201. doi.org/10.1103/PhysRevLett.120.163201.

Willke, P., T. Bilgeri, X. Zhang, Y. Wang, C. Wolf, H. Aubin, A. Heinrich, and T. Choi. 2021. “Coherent Spin Control of Single Molecules on a Surface.” ACS Nano 15(11):17959–17965. doi.org/10.1021/Acsnano.1c06394.

Wilson, A., J. Lawrence, E. C. Yang, M. Nakano, D. N. Hendrickson, and S. Hill. 2006. “Magnetization Tunneling in High-Symmetry Single-Molecule Magnets: Limitations of the Giant Spin Approximation.” Physical Review B 74(14):140403. doi.org/10.1103/Physrevb.74.140403.

Wiltschko, R., and W. Wiltschko. 2012. “Magnetoreception.” In Sensing in Nature, edited by C. López-Larrea, 126–141. New York: Springer US.

Wiltschko, W., and R. Wiltschko. 2005. “Magnetic Orientation and Magnetoreception in Birds and Other Animals.” Journal of Comparative Physiology A 191(8):675–693. doi.org/10.1007/S00359-005-0627-7.

Wolf, F., Y. Wan, J. C. Heip, F. Gebert, C. Shi, and P. O. Schmidt. 2016. “Non-Destructive State Detection for Quantum Logic Spectroscopy of Molecular Ions.” Nature 530(7591):457–460. doi.org/10.1038/Nature16513.

Wolfowicz, G., and J. J. L. Morton. 2016. “Pulse Techniques for Quantum Information Processing.” eMagRes 5:1515–1528.

Wong, L. J., and I. Kaminer. 2021. “Prospects in X-Ray Science Emerging from Quantum Optics and Nanomaterials.” Applied Physics Letters 119(13):130502. doi.org/10.1063/5.0060552.

Wrachtrup, J., C. Von Borczyskowski, J. Bernard, M. Orrit, and R. Brown. 1993. “Optical Detection of Magnetic Resonance in a Single Molecule.” Nature 363(6426):244–245. doi.org/10.1038/363244a0.

Wright, K., K. M. Beck, S. Debnath, J. M. Amini, Y. Nam, N. Grzesiak, J. S. Chen, N. C. Pisenti, M. Chmielewski, C. Collins, K. M. Hudek, J. Mizrahi, J. D. Wong-Campos, S. Allen, J. Apisdorf, P. Solomon, M. Williams, A. M. Ducore, A. Blinov, S. M. Kreikemeier, V. Chaplin, M. Keesan, C. Monroe, and J. Kim. 2019. “Benchmarking an 11-Qubit Quantum Computer.” Nature Communications 10(1):5464. doi.org/10.1038/S41467-019-13534-2.

Yan, R., Y. Zhang, Y. Li, L. Xia, Y. Guo, and Q. Zhou. 2020. “Structural Basis for the Recognition of SARS-CoV-2 by Full-Length Human ACE2.” Science 367(6485):1444–1448. doi.org/10.1126/science.abb2762.

Yang, K., W. Paul, S.-H. Phark, P. Willke, Y. Bae, T. Choi, T. Esat, A. Ardavan, A. J. Heinrich, and C. P. Lutz. 2019. “Coherent Spin Manipulation of Individual Atoms on a Surface.” Science 366(6464):509–512. doi.org/10.1126/Science.Aay6779.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×

Ye, L., and S. Mukamel. 2020. “Interferometric Two-Photon-Absorption Spectroscopy with Three Entangled Photons.” Applied Physics Letters 116(17):174003. doi.org/10.1063/5.0004617.

Young, L., K. Ueda, M. Gühr, P. H. Bucksbaum, M. Simon, S. Mukamel, N. Rohringer, K. C. Prince, C. Masciovecchio, M. Meyer, A. Rudenko, D. Rolles, C. Bostedt, M. Fuchs, D. A. Reis, R. Santra, H. Kapteyn, M. Murnane, H. Ibrahim, F. Légaré, M. Vrakking, M. Isinger, D. Kroon, M. Gisselbrecht, A. L’Huillier, H. J. Wörner, and S. R. Leone. 2018. “Roadmap of Ultrafast X-Ray Atomic and Molecular Physics.” Journal of Physics B: Atomic, Molecular, and Optical Physics 51(3):032003. doi.org/10.1088/1361-6455/Aa9735.

Yu, C.-J., M. J. Graham, J. M. Zadrozny, J. Niklas, M. D. Krzyaniak, M. R. Wasielewski, O. G. Poluektov, and D. E. Freedman. 2016. “Long Coherence Times in Nuclear Spin-Free Vanadyl Qubits.” Journal of the American Chemical Society 138(44):14678–14685. doi.org/10.1021/Jacs.6b08467.

Yu, C.-J., S. von Kugelgen, D. W. Laorenza, and D. E. Freedman. 2021. “A Molecular Approach to Quantum Sensing.” ACS Central Science 7(5):712–723. doi.org/10.1021/Acscentsci.0c00737.

Zadrozny, J. M., A. T. Gallagher, T. D. Harris, and D. E. Freedman. 2017. “A Porous Array of Clock Qubits.” Journal of the American Chemical Society 139(20):7089–7094. doi.org/10.1021/Jacs.7b03123.

Zadrozny, J. M., J. Liu, N. A. Piro, C. J. Chang, S. Hill, and J. R. Long. 2012. “Slow Magnetic Relaxation in a Pseudotetrahedral Cobalt(II) Complex with Easy-Plane Anisotropy.” Chemical Communications 48(33):3927–3929. doi.org/10.1039/C2CC16430B.

Zadrozny, J. M., J. Niklas, O. G. Poluektov, and D. E. Freedman. 2015. “Millisecond Coherence Time in a Tunable Molecular Electronic Spin Qubit.” ACS Central Science 1(9):488–492. doi.org/10.1021/Acscentsci.5b00338.

Zadrozny, J. M., D. J. Xiao, J. R. Long, M. Atanasov, F. Neese, F. Grandjean, and G. J. Long. 2013. “Mössbauer Spectroscopy as a Probe of Magnetization Dynamics in the Linear Iron(I) and Iron(II) Complexes [Fe(C(SiMe3)3)2]1-/0.” Inorganic Chemistry 52(22):13123–13131. doi.org/10.1021/Ic402013n.

Zhang, X., C. Wolf, Y. Wang, H. Aubin, T. Bilgeri, P. Willke, A. J. Heinrich, and T. Choi. 2022. “Electron Spin Resonance of Single Iron Phthalocyanine Molecules and Role of Their Non-Localized Spins in Magnetic Interactions.” Nature Chemistry 14(1):59–65. doi.org/10.1038/S41557-021-00827-7.

Zhu, G.-Z., D. Mitra, B. L. Augenbraun, C. E. Dickerson, M. J. Frim, G. Lao, Z. D. Lasner, A. N. Alexandrova, W. C. Campbell, J. R. Caram, J. M. Doyle, and E. R. Hudson. 2022. “Functionalizing Aromatic Compounds with Optical Cycling Centres.” Nature Chemistry 14(9):995–999. doi.org/10.1038/S41557-022-00998-X.

Zoltowski, B. D., Y. Chelliah, A. Wickramaratne, L. Jarocha, N. Karki, W. Xu, H. Mouritsen, P. J. Hore, R. E. Hibbs, C. B. Green, and J. S. Takahashi. 2019. “Chemical and Structural Analysis of a Photoactive Vertebrate Cryptochrome from Pigeon.” Proceedings of the National Academy of Sciences USA 116(39):19449–19457. doi.org/10.1073/pnas.1907875116.

Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 63
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 64
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 65
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 66
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 67
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 68
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 69
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 70
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 71
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 72
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 73
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 74
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 75
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 76
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 77
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 78
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 79
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 80
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 81
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 82
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 83
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 84
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 85
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 86
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 87
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 88
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 89
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 90
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 91
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 92
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 93
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 94
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 95
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 96
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 97
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 98
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 99
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 100
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 101
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 102
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 103
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 104
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 105
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 106
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 107
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 108
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 109
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 110
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 111
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 112
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 113
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 114
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 115
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 116
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 117
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 118
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 119
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 120
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 121
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 122
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 123
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 124
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 125
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 126
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 127
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 128
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 129
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 130
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 131
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 132
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 133
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 134
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 135
Suggested Citation:"3 Measurement and Control of Molecular Quantum Systems." National Academies of Sciences, Engineering, and Medicine. 2023. Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States. Washington, DC: The National Academies Press. doi: 10.17226/26850.
×
Page 136
Next: 4 Experimental and Computational Approaches for Scaling Qubit Design and Function »
Advancing Chemistry and Quantum Information Science: An Assessment of Research Opportunities at the Interface of Chemistry and Quantum Information Science in the United States Get This Book
×
Buy Paperback | $40.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

The field of quantum information science (QIS) has witnessed a dramatic rise in scientific research activities in the 21st century as excitement has grown about its potential to revolutionize communications and computing, strengthen encryption, and enhance quantum sensing, among other applications. While, historically, QIS research has been dominated by the field of physics and computer engineering, this report explores how chemistry - in particular the use of molecular qubits - could advance QIS. In turn, researchers are also examining how QIS could be used to solve problems in chemistry, for example, to facilitate new drug and material designs, health and environmental monitoring tools, and more sustainable energy production.

Recognizing that QIS could be a disruptive technology with the potential to create groundbreaking products and new industries, Advancing Chemistry and Quantum Information Science calls for U.S. leadership to build a robust enterprise to facilitate and support research at the intersection of chemistry and QIS. This report identifies three key research areas: design and synthesis of molecular qubit systems, measurement and control of molecular quantum systems, and experimental and computational approaches for scaling qubit design and function. Advancing Chemistry and Quantum Information Science recommends that the Department of Energy, National Science Foundation, and other funding agencies should support multidisciplinary and collaborative research in QIS, the development of new instrumentation, and facilities, centralized and open-access databases, and efforts to create a more diverse and inclusive chemical workforce.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    Switch between the Original Pages, where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

    « Back Next »
  6. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  7. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  8. ×

    View our suggested citation for this chapter.

    « Back Next »
  9. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!