Navigation and positioning technologies play a key role in modern human activities. Currently, navigation devices are mostly dependent on the global navigation satellite systems, for example, the global positioning system (GPS). However, it is known that GPS signals can be jammed or spoofed or otherwise fail. As the need for GPS independent navigation increases in some practical application scenarios, novel navigation techniques based on the geomagnetic field have been developed. A main task of geomagnetic navigation is to obtain the amplitude and direction of the geomagnetic field accurately. Here, we introduce an alternative scheme for vectorial measurements of the local geomagnetic field for magnetic positioning based on the biological ferric sulfide cluster, which exists in the magnetoreceptor protein/cryptochrome complex in certain avian species. We find that by observing the number of peaks and the proportional rate of spectrum on resonance, both the direction and intensity of the magnetic field can be determined. Therefore, our findings may provide a fresh insight into magnetic field measurement and also suggest further guidelines for the design and operation of satellite-free navigation systems based on the electrically tunable inorganic biological molecules.

The geomagnetic field is an inherent resource of Earth, which has a wide variety of applications, such as in earth science, navigation, mineral exploration, earthquake prediction, and other fields.1,2 Earth’s magnetic field is an omnipresent three-dimensional vector field, which has one-to-one correspondence with the latitude and longitude, providing a natural coordinate system for navigation in GPS-denied environments. Therefore, in principle, as long as the geomagnetic field vector of the geographical point on Earth is measured, global positioning can be realized.3 

Remarkably, long-distance migratory animals appear to have such a natural global positioning capability.4,5 Compelling behavioral evidence demonstrated that migratory birds can potentially use the information from Earth’s magnetic field for orientational and navigational purposes.6–9 The birds utilize the direction of the geomagnetic field as a magnetic compass10 to maintain a directed heading while sensing the intensity of the geomagnetic field as a component of the navigational magnetic map11,12 to assess their approximate geographic position. However, the underlying mechanism of how birds use the geomagnetic field to navigate remains an enduring mystery.13,14 One leading hypothesis assumes that the radical pair mechanism15–24 might account for the avian detection of the inclination of the geomagnetic field. Moreover, the magnetic map sensitivity of birds is likely to rely on biogenic magnetite25–28 associated with trigeminal nerves. A novel biocompass model29 based on the magnetoreceptor protein (MagR)/cryptochrome (Cry) complex has been suggested to explain the avian magnetoreception. Recently, we also proposed a compass-free migratory navigation hypothesis30 based on the spectrum of resonance fluorescence of a four-level model derived from the ferric sulfide cluster, which exists in the MagR/Cry protein complex. We have shown that the feature of the spectrum can be a cue for animal navigation. We also showed that the weak radio frequency magnetic field can disrupt avian magnetic orientation.

Meanwhile, for engineered systems, the geomagnetic field can be used in global positioning and navigation that do not rely on GPS satellites. The main objective of making geomagnetic navigation possible is achieving an accurate measurement of geomagnetic field information. Several different styles of magnetic sensors31 have been developed to detect the presence, strength, or direction of magnetic fields. One well-established magnetic sensor is the superconducting quantum interference device (SQUID),32–34 which is a scalar-type magnetic sensor. SQUIDs have been proved to be reliable mechanical geomagnetic surveying and mapping equipment. This device has the advantages of a wide measurement range and high sensitivity on the order of 1 fT/Hz, but the disadvantage is that it requires cryogenic cooling. The insulation and refrigeration equipment leads to the magnetometer’s large volume, high cost, and inflexible installation.35–37 Another class of sensitive magnetic sensor is the optically pumped magnetometer (OPM),38–40 which is based on the Zeeman effect in alkali atoms. As for OPMs, the atomic vapor is encapsulated in the glass bubble, and the magnetic resonance signal is extracted by the interaction between atoms and the environmental magnetic field so as to obtain the intensity of the magnetic field. OPMs are known as precise sensors for scalar measurement of Earth’s magnetic field without the hurdle of cryogenic operation, but they suffer from so-called dead zones41 and heading errors42,43 when used for modern geomagnetic surveys.

The aforementioned studies motivate us to perceive an alternative method to measure Earth’s magnetic field for global positioning, which is also inspired by our four-level system derived from the ferric sulfide cluster.29,30 By studying the resonance fluorescence spectrum of the system, the direction of the geomagnetic field can be determined by tuning the applied artificial electric field and observing the characteristics of the spectrum. The resonance fluorescence spectrum always exhibits three peaks as long as the direction of the geomagnetic field is parallel to the electric field. Moreover, the magnitude of the geomagnetic field can be evaluated by the proportional rate of observing the spectrum on resonance. From the viewpoint of spin dynamics, we numerically inspect the spin excitations of the cluster system and clarify that the feature of the spin resonance spectra also provides the precise information of the direction of the magnetic field.

We shall specifically refer to the MagR/Cry complex,29 a potential magnetoreceptor that consists of ferric sulfide cluster protein. As schematically depicted in Fig. 1(a), we thus devised the ferric sulfide cluster as a three-site cluster minimal model consisting of two magnetic-ion sites (M) and a ligand site (L). Our simplified M-L-M cluster system under both an external (geo-)magnetic field B and an (bio-)electric field E is described by the following Hamiltonian:30 

(1)

The first term in Eq. (1) refers to the contribution of the Zeeman energy of two spins S1 and S2 in the applied magnetic field, and g and μB, respectively, denote the g-factor and the Bohr magneton, while the second term arises from the magnetoelectric coupling between the spin-induced electric polarization and the applied electric field. The non-collinearity of two neighboring spins in the M-L-M cluster brings in a local electric polarization P via the spin-current mechanism44 of the following form:

(2)

where gME is the magnetoelectric coupling that depends on the spin–orbit interaction and e12=sinθcosϕ,sinθsinϕ,cosθ denotes a unit vector connecting the neighboring spins S1 and S2 in an arbitrary direction (θ, ϕ) relative to the coordinates (ξ, η, ζ), as shown in Fig. 1(b). For a Cartesian coordinate ξ-η-ζ, any vector is expressed as V = (Vξ, Vη, Vζ). We choose the direction of the magnetic field B as the ζ-axis and assume that there is a tilt angle ϑ between B and the electric field E, which can be written as E = E(sin ϑ, 0, cos ϑ) without loss of generality.

FIG. 1.

(a) Schematic picture of the considered three-site M-L-M cluster model, in which M and L indicate a magnetic ion and a ligand ion, respectively. The arrows on the two magnetic sites denote two noncollinear spins S1 and S2. The spin-driven polarization relevant to the cluster obeys the relation Pe12 × (S1 × S2), where e12=sinθcosϕ,sinθsinϕ,cosθ is the unit vector pointing from S1 to S2. (b) The geometry of the (ξ, η, ζ) coordinate frame with the ζ-axis parallel to the magnetic field B direction; the electric field E = E(sin ϑ, 0, cos ϑ) has a tilt angle ϑ toward B. The η-axis is perpendicular to the plane containing B and E, and the ξ-axis is chosen to satisfy eξ=eη×eζ.

FIG. 1.

(a) Schematic picture of the considered three-site M-L-M cluster model, in which M and L indicate a magnetic ion and a ligand ion, respectively. The arrows on the two magnetic sites denote two noncollinear spins S1 and S2. The spin-driven polarization relevant to the cluster obeys the relation Pe12 × (S1 × S2), where e12=sinθcosϕ,sinθsinϕ,cosθ is the unit vector pointing from S1 to S2. (b) The geometry of the (ξ, η, ζ) coordinate frame with the ζ-axis parallel to the magnetic field B direction; the electric field E = E(sin ϑ, 0, cos ϑ) has a tilt angle ϑ toward B. The η-axis is perpendicular to the plane containing B and E, and the ξ-axis is chosen to satisfy eξ=eη×eζ.

Close modal

To inspect the spin dynamics of the M-L-M cluster, it is convenient to use the phenomenological Landau–Lifshitz–Gilbert (LLG) equation, which is given by45 

(3)

where αG (=0.01) is the phenomenological Gilbert-damping coefficient and Hieff=H/Si is the effective magnetic field that acts on the ith spin Si. Each spin is normalized to be of unit length (i.e., S = 1), and the LLG equation is solved numerically with the fourth-order Runge–Kutta method.

We calculate the spin excitation modes and their resonance frequencies through the dynamical magnetic susceptibility,45 which is defined as

(4)

where the subscript μ stands for the orthogonal axes ξ, η, or ζ, and the ζ-axis is defined to be parallel to the static external magnetic field B. Here, Bμ(ω) and ΔSμ(ω) are Fourier transform of the magnetic field pulse Bμ(t) and that of the simulated time-profile of the spatially averaged spin ΔS(t) = S(t) − S(0), with S(t) = (1/N)iSi(t), which is the transient response of the system under a magnetic field pulse. In our numerical simulation, we adopt an intensive short pulse with a δ-function shape, i.e., B(t)=δ(t)eξ applied at t = 0. We initialize the cluster system with a random spin configuration and relax it for a sufficient time to achieve a ground state by using the LLG equation simulation. After stabilizing the system, we then simulate the spin dynamics by applying an intensive pulse of magnetic field to obtain the spin excitation spectra.

We first calculate the spin excitation spectra of the model system and discuss how the tilt angle of the applied electric field E and the orientation of the unit vector e12 affect the characteristics of the spectral peaks. We show in Figs. 2(a)2(d) the calculated imaginary parts of the dynamical magnetic susceptibility Imχξ(ω) of the M-L-M cluster under tilted electric fields (i.e., ϑ = π/4) for various selected orientations of the unit vector e12 with (θ, ϕ) = (π/2, π/2), (π/4, π/2), (π/3, 0.1843π), and (2π/3, 0.6959π), which correspond to the relative angle between E and e12 as π/2, π/3, π/6, and 3π/4, respectively. Here, we fix E = 0.06 and B = 0.1 for the calculations. Note that the spin excitation modes are seen as peaks in the spectrum of Imχξ(ω). We find that each of the spectra has two distinct peaks, indicating the existence of two spin resonance modes for all orientations of e12. For comparison, Figs. 2(e)2(h) show the calculated spectra Imχξ(ω) of the M-L-M cluster under a parallel electric field (i.e., ϑ = 0 with EB) for various selected orientations e12. The spectra of Imχξ(ω) clearly exhibit a single resonance peak for all e12 orientations. These facts indicate that the peak numbers of the spectra Imχξ(ω) involve information of the relative direction of the magnetic field B and the electric field E. When the applied electric field E has a tilt angle with respect to the applied magnetic field B, the spectra of Imχξ(ω) exhibit two distinct resonance peaks for a randomly oriented e12; otherwise, the spectra exhibit a single resonance peak, which indicates the magnetic field B and electric field E are parallel to each other. This property may allow a navigational signal for the birds to perform precise geomagnetic field direction sensing by adjusting the bioelectric field direction.

FIG. 2.

Calculated imaginary part of the dynamical magnetic susceptibility Imχξ(ω). The calculations are carried out for various orientations e12=sinθcosϕ,sinθsinϕ,cosθ with different angles θ and ϕ in different panels. In panels (a) to (d), the electric field is applied with an inclination angle of π/4 with respect to the magnetic field B, i.e., E=(E/2,0,E/2), while in panels (e) to (h), it is parallel to B. Those data are calculated for a fixed E = 0.06 and B = 0.1. Here, the frequency ω is measured in units of ωL = BB with g being the g-factor and μB being the Bohr magneton.

FIG. 2.

Calculated imaginary part of the dynamical magnetic susceptibility Imχξ(ω). The calculations are carried out for various orientations e12=sinθcosϕ,sinθsinϕ,cosθ with different angles θ and ϕ in different panels. In panels (a) to (d), the electric field is applied with an inclination angle of π/4 with respect to the magnetic field B, i.e., E=(E/2,0,E/2), while in panels (e) to (h), it is parallel to B. Those data are calculated for a fixed E = 0.06 and B = 0.1. Here, the frequency ω is measured in units of ωL = BB with g being the g-factor and μB being the Bohr magneton.

Close modal

To further elucidate the influence of the tilted angle ϑ of E on the spin excitation of the M-L-M cluster, we show in Fig. 3(a) the spectra Imχξ(ω) for various cases ϑ. In the simulation, the strength of the magnetic and electric field is fixed at B = 0.1 and E = 0.06, while the direction of the unit vector e12 = (0, 1, 0) is along the η-axis. We find that as ϑ decreases from π/2 to 0, the resonance frequency of the lower-lying peak (ω1) shifts to high values while that of the higher-lying peak (ω2) barely changes. However, the intensity of the higher-lying peak ω2 decreases with decreasing ϑ, and it finally disappears at ϑ = 0. The plot in the inset of Fig. 3(a) illustrates how the difference in the resonance frequencies Δω (=ω2ω1) between the higher-lying peak and lower-lying peak varies as a function of the tilted angle ϑ. It shows that as ϑ decreases, Δω monotonically decreases. Therefore, the peak’s difference Δω would be sensitive to the relative direction of B and E, which can be a cue for birds to adjust the angle of E and then sense the direction of the magnetic field B. In Fig. 3(b), we show Imχξ(ω) when the magnetic field and the electric field are parallel to each other, i.e., BEeζ, for several values of B with fixed E = 0.06. It can be found from Fig. 3(b) that the resonance frequencies are significantly influenced by the amplitude of B. We can find that the positions of resonance peaks increase whereas the intensities of the peaks are monotonically decreased as B increases. The frequencies of the resonance modes (ωR) increase nearly linearly with the magnitude of B, as illustrated in the inset of Fig. 3(b).

FIG. 3.

(a) Simulated imaginary part of the dynamical magnetic susceptibility Imχξ(ω) for several values of ϑ with B = 0.1, E = 0.06 and the direction of e12 along the η-axis. The inset shows the difference in resonance frequencies Δω(=ω2ω1) between the higher-lying peak (ω2) and the lower-lying peak (ω1) as a function of ϑ. (b) The imaginary part of the dynamical magnetic susceptibility Imχξ(ω) calculated by keeping the electric and magnetic fields in parallel, i.e., BEeζ, while varying the value B of the external magnetic field. The inset shows B’s dependence on the resonance frequencies ωR [corresponding to the peak position of the spectra Imχξ(ω)]. The frequency ω is measured in units of ωL = BB with B = 0.1.

FIG. 3.

(a) Simulated imaginary part of the dynamical magnetic susceptibility Imχξ(ω) for several values of ϑ with B = 0.1, E = 0.06 and the direction of e12 along the η-axis. The inset shows the difference in resonance frequencies Δω(=ω2ω1) between the higher-lying peak (ω2) and the lower-lying peak (ω1) as a function of ϑ. (b) The imaginary part of the dynamical magnetic susceptibility Imχξ(ω) calculated by keeping the electric and magnetic fields in parallel, i.e., BEeζ, while varying the value B of the external magnetic field. The inset shows B’s dependence on the resonance frequencies ωR [corresponding to the peak position of the spectra Imχξ(ω)]. The frequency ω is measured in units of ωL = BB with B = 0.1.

Close modal

As a step to calculate the resonance fluorescence spectra, we should solve the eigenenergies and eigenstates of the model system. Following our previous paper,30 we restate the derivation of a four-level system from the model Hamiltonian (1). Three operators are conveniently defined, namely, the total spin operator S = (S1 + S2), a new operator Q = 2S1 × S2, and a vector field C = (E × e12)/2, which satisfy the commutation relations as follows:

where the subscripts i, j, k = ξ, η, ζ. Their further combinations J = (S + Q)/2 and K = (SQ)/2 give rise to two sets of decoupled SU(2) generators. Then one can rewrite the model Hamiltonian (1) as

(5)

Clearly, it indicates an SU(2) × SU(2) model characterized by two vector fields given by α = (B + C) and β = (BC), which can be tuned by both the magnetic and electric fields.

If one chooses the direction of α as the ξ-axis of our coordinate frame, the SU(2) × SU(2) Hamiltonian (5) can be simplified as

(6)

which is solved by four eigenvectors: the ground sate |g⟩ and three excited states |a⟩, |b⟩, and |c⟩, with the corresponding eigenenergies Eg=(α+β)/2, Ea=(αβ)/2, Eb=(βα)/2, and Ec=(α+β)/2, respectively. Therefore, one can construct a four-level system parameterized by α and β, and the corresponding Hamiltonian (5) can be written in Fock space as

(7)

The energy-level scheme of the four-level system is shown in Fig. 4(d), which is employed as the model system to calculate the resonance fluorescence spectrum.

FIG. 4.

(a) Three-dimensional plots of the resonance fluorescence spectra as functions of ω and the difference between α and β, i.e., βα, which varies from −10Γ to 10Γ by keeping α/Γ = 20. For βα = 0, the spectra exhibits a three-peak structure; the others exhibit a five-peak structure. (b) The resonance fluorescence spectra vs ω and φ, where φ is the angle between B and C ranging from π/8 to 7π/8, and the corresponding Larmor frequencies caused by B and C are set to 3Γ and 10Γ, respectively. For φ = π/2, the spectrum exhibits a three-peak structure, while others exhibit a five-peak structure. (c) The resonance fluorescence spectra vs ω and the magnetic field B for the case of φ = π/2. The corresponding Larmor frequency caused by B varies from 2Γ to 24Γ while the direction of B is perpendicular to that of C. (d) Energy-level diagram of the derived four-level model showing the allowed transitions among the energy-levels. α and β are the spacing of the corresponding energy levels.

FIG. 4.

(a) Three-dimensional plots of the resonance fluorescence spectra as functions of ω and the difference between α and β, i.e., βα, which varies from −10Γ to 10Γ by keeping α/Γ = 20. For βα = 0, the spectra exhibits a three-peak structure; the others exhibit a five-peak structure. (b) The resonance fluorescence spectra vs ω and φ, where φ is the angle between B and C ranging from π/8 to 7π/8, and the corresponding Larmor frequencies caused by B and C are set to 3Γ and 10Γ, respectively. For φ = π/2, the spectrum exhibits a three-peak structure, while others exhibit a five-peak structure. (c) The resonance fluorescence spectra vs ω and the magnetic field B for the case of φ = π/2. The corresponding Larmor frequency caused by B varies from 2Γ to 24Γ while the direction of B is perpendicular to that of C. (d) Energy-level diagram of the derived four-level model showing the allowed transitions among the energy-levels. α and β are the spacing of the corresponding energy levels.

Close modal

Following the standard steps,46,47 we can derive the resonance fluorescence spectrum function of the four-level system as follows:30 

(8)

where P(+)(t)=(P()(t)) is the effective polarization operator with P(+) (t) = d1γ1(t) + d2γ2(t). Here, d1 and d2 denote the dipole moments, γ1 = |g⟩⟨a| + |b⟩⟨c| and γ2 = |g⟩⟨b| + |a⟩⟨c| are the energy-lowering operators related to the four-level systems. Generally, it is difficult to derive an exact analytical expression for the spectrum S(ω). Nevertheless, we found that it is possible to solve an analytical spectrum for our four-level system in the regime α = β under the double-resonance conditions.30 Now, we take a close look at B and C, assuming that the angle between B and C is φ. Then the expressions for α and β can be expressed concretely as

(9)

With Eq. (9) in mind, we then investigate the double resonance condition α = β = ν. From the above-mentioned equation, we can see that α and β are equal to each other when cos φ = 0, i.e., φ = π/2. Furthermore, we recall that

where ψ denotes the angle between E and e12. Thus, the final formula for the double resonance condition α = β = ν in terms of B and E can be written as

(10)

Here, ψ0 specifies the angle when the double resonance condition is fulfilled. In the following, we would calculate the spectrum numerically for various parameters in more general situations.

Figure 4 shows the numerical results for the three-dimensional plots of resonance fluorescence spectra of our four-level system under various situations. First, Fig. 4(a) shows the spectrum vs ω and the difference between β and α, i.e., βα, which varies from −10Γ to 10Γ (with Γ being the natural width) by keeping α/Γ = 20. Here, we choose Ω/Γ = 10 and ν/Γ = 30, where Ω stands for the Rabi frequency associated with the driving field of frequency ν. It is seen that when αβ, the spectrum generally exhibits a characteristic five-peaked resonance structure, where there is a central peak and two pairs of side-peaks symmetrically located at the left- and right-hand side, respectively. When βα approaches zero, the magnitude of these side-peaks grows, and the separation between the inner and outer side-peak is getting smaller. It is clearly shown that the five-peaked spectrum collapses into a three-peaked structure at βα = 0, which actually serves as a three-level system.30 

In Fig. 4(b), we present the resonance fluorescence spectra as a function of both ω and φ, where φ is the angle between B and C. In this situation, the magnitude of α and β is no longer addition and subtraction of |B| and |C|, as we simplified previously.30 Instead, they are replaced by Eq. (9). We choose φ ranging from π/8 to 7π/8, while the magnitude of |B| and |C| is set to 3Γ and 10Γ, respectively. One finds that for the given amplitude of B and C, the spectrum is triple peaked for φ = π/2, in which B is perpendicular to C. In this case, one central peak is located at ω = ν while two side-peaks are symmetrically distributed to both sides. However, as φ deviates from π/2, the two side-peaks will split into two peaks, and a five-peak feature appears. We can see that the more difference there is between the angle φ and π/2, the bigger the splitting of the two side peaks in the spectrum is.

Finally, we consider a special situation, φ = π/2. In Fig. 4(c), we show the resonance fluorescence spectra as a function of ω and the magnitude of B under the condition that B is perpendicular to C. The Larmor frequency caused by C is kept at 4Γ, and the other parameters are set to Ω/Γ = 6 and ν/Γ = 30. One finds that the spectrum generally exhibits a three-peak structure when B is perpendicular to C. The intensity of the central and side peaks grows when the magnetic field B increases. Furthermore, the spacing of the side-peaks from the central peak decreases as B increases.

We can clearly determine the direction of the geomagnetic field from the characteristics of the resonance fluorescence spectra described in Fig. 4. As shown in Figs. 4(b) and 4(c), the spectrum will always exhibit a three-peak structure when B is perpendicular to C, i.e., φ = π/2; otherwise, the spectrum exhibits a five-peak structure. As we know, parameter C is defined as C = (E × e12)/2, and the direction of e12 is chosen randomly in our model; thus, all possible values of C lie in a plane perpendicular to E. Therefore, if the spectrum generally exhibits a three-peak structure, it means that B is perpendicular to C; thus, B is exactly parallel to the direction of E in this case. This implies that the spectrum hosting a three-peak structure is a characteristic signature of the parallel nature of the electric field E and magnetic field B.

In Subsection II E, we have determined the direction of the magnetic field by the direction of E, and we will get the magnitude of the magnetic field by the proportional rate of resonance here.

We assume the angle between E and e12 is ψ; thus, all possible values of e12 constitute a sphere. According to the resonant condition illustrated in Eq. (10), it is obvious to see that the spectrum will be resonant only in the case of ψ = ψ0 or ψ = πψ0, and this corresponds to two circular lines on the sphere. In order to get a finite area, we relax the resonant condition to ψ0ψ0 +  and π − (ψ0 + ) ∼ πψ0. The area of two bands is dS = 4π sin ψ0. The proportional rate of observing the resonance spectrum is

(11)

In the following, we will make an estimation of . We assume that the spectrum for ψ0 + is slightly different from that of ψ0 by nearly about Γ, where Γ is the half width of the spectrum. According to the Rayleigh criterion, one cannot distinguish two peaks if their distance is within the half width. Moreover, we consider that the spectrum for ψ0 + is resonant with ν + Γ. Recalling the resonant condition in Eq. (10), we have

(12a)
(12b)

Subtracting Eq. (12a) from Eq. (12b) and after some simplification, we can get

(13)

In deriving the above-mentioned formula, we have made the assumption that is a small quantity.

Substituting Eq. (13) into Eq. (11) and moving cos ψ0 to the left hand of the equation, we will finally get the expression

(14)

Once we know cos ψ0, we can substitute Eq. (14) into Eq. (12a) and get the explicit expression of |B| caused by the magnetic field, which is given by

(15)

where Γ is the half width of the resonance fluorescence spectrum. The above-mentioned explicit expression indicates that the magnitude of the magnetic field is determined by the parameter P, the proportional rate of observing the spectrum of resonance fluorescence, which is related to the angle ψ. Therefore, it allows one to deduce the magnetic field intensity by measuring the proportional rate for detection of the resonance fluorescence in the experiments.

In summary, we have investigated a four-level system derived from the biological ferric sulfide cluster existing in the MagR/Cry protein complex. By making Laplace transform of the double-time correlation function of the electric field and utilizing the quantum regression theorem, the resonance fluorescence spectrum of this four-level system is obtained. Further analysis reveals that the spectrum generally exhibits a five-peak structure, with one central peak located at the driving frequency ν and four side peaks. However, when α = β, the spectrum will contain only three peaks, with one central peak and two side peaks symmetrically distributed at the left- and right-hand side, respectively. Moreover, the separation between the side peaks and the central peak grows when α or β is farther away from resonance, with the driving frequency ν increasing. Thus, a new method for measuring Earth’s magnetic field is proposed based on the characteristics of the spectrum. Explicitly, the direction of the geomagnetic field can be determined by tuning the applied electric field. When the spectrum is always observed as three peaks, the direction of the magnetic field is the same with the applied electric field. In addition, the magnitude of the geomagnetic field is related to the proportional rate of observing the spectrum on resonance. Our study provides an entirely new insight into magnetic field measurement and may be helpful for potential manmade practical devices.

We should note that the iron-sulfur cluster in the intact MagR protein binds either the 2Fe–2S or 3Fe–4S form.48–50 It is reasonable to extend our model calculations to the 2Fe–2S cluster since the spin-induced polarization in the 2M-2L cluster exhibits an analogous form to Eq. (2) with an anisotropic amplitude.51 Extensions of this work to more realistic 2Fe–2S or 3Fe–4S cluster situations are future works of interest.

This work was supported by the National Key R&D Program of China, Grant No. 2017YFA0304304, and the NSFC, Grant No. 11935012.

The authors have no conflicts to disclose.

Xin Zhao: Investigation (equal); Methodology (equal); Writing – original draft (equal). Hong-Bo Chen: Investigation (equal); Methodology (equal); Writing – original draft (equal). Li-Hua Lu: Formal analysis (lead); Project administration (lead); Writing – review & editing (equal). You-Quan Li: Conceptualization (lead); Supervision (lead); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
F.
Goldenberg
, “
Geomagnetic navigation beyond the magnetic compass
” in
2006 IEEE/ION Position, Location, and Navigation Symposium (PLANS)
, (
IEEE
,
2006
), pp.
684
694
.
2.
M. N.
Nabighian
,
V. J. S.
Grauch
,
R. O.
Hansen
,
T. R.
Lafehr
,
Y.
Li
,
J. W.
Peirce
,
J. D.
Phillips
, and
M. E.
Ruder
, “
The historical development of the magnetic method in exploration
,”
Geophysics
70
,
33ND
61ND
(
2005
).
3.
A.
Canciani
and
J.
Raquet
, “
Absolute positioning using the Earth’s magnetic anomaly field
,”
Navigation
63
,
111
126
(
2016
).
4.
R.
Wiltschko
and
W.
Wiltschko
,
Magnetic Orientation in Animals
(
Springer-Verlag
,
Berlin
,
1995
).
5.
K. J.
Lohmann
,
C. M. F.
Lohmann
, and
N. F.
Putman
, “
Magnetic maps in animals: nature’s GPS
,”
J. Exp. Biol.
210
,
3697
3705
(
2007
).
6.
W.
Wiltschko
and
F.
Merkel
, “
Orientierung zugunruhiger Rotkehlchen im statischen Magnetfeld
,”
Verh. der Dtsch. Zool. Ges.
59
,
362
367
(
1966
).
7.
T.
Ritz
,
P.
Thalau
,
J. B.
Phillips
,
R.
Wiltschko
, and
W.
Wiltschko
, “
Resonance effects indicate a radical-pair mechanism for avian magnetic compass
,”
Nature
429
,
177
180
(
2004
).
8.
W.
Wiltschko
and
R.
Wiltschko
, “
Magnetic orientation and magnetoreception in birds and other animals
,”
J. Comp. Physiol., A
191
,
675
693
(
2005
).
9.
J.
Wynn
,
O.
Padget
,
H.
Mouritsen
,
J.
Morford
,
P.
Jaggers
, and
T.
Guilford
, “
Magnetic stop signs signal a European songbird’s arrival at the breeding site after migration
,”
Science
375
,
446
449
(
2022
).
10.
W.
Wiltschko
and
R.
Wiltschko
, “
Magnetic compass of European robins
,”
Science
176
,
62
64
(
1972
).
11.
J. L.
Gould
, “
The map sense of pigeons
,”
Nature
296
,
205
211
(
1982
).
12.
K. J.
Lohmann
,
K. M.
Goforth
,
A. G.
Mackiewicz
,
D. S.
Lim
, and
C. M. F.
Lohmann
, “
Magnetic maps in animal navigation
,”
J. Comp. Physiol., A
208
,
41
67
(
2022
).
13.
H.
Mouritsen
, “
Long-distance navigation and magnetoreception in migratory animals
,”
Nature
558
,
50
59
(
2018
).
14.
C.
Xie
, “
Searching for unity in diversity of animal magnetoreception: From biology to quantum mechanics and back
,”
Innovation
3
,
100229
(
2022
).
15.
K.
Schulten
,
C. E.
Swenberg
, and
A.
Weller
, “
A biomagnetic sensory mechanism based on magnetic field modulated coherent electron spin motion
,”
Z. Phys. Chem.
111
,
1
5
(
1978
).
16.
T.
Ritz
,
S.
Adem
, and
K.
Schulten
, “
A model for photoreceptor-based magnetoreception in birds
,”
Biophys. J.
78
,
707
718
(
2000
).
17.
C. T.
Rodgers
and
P. J.
Hore
, “
Chemical magnetoreception in birds: The radical pair mechanism
,”
Proc. Natl. Acad. Sci. U. S. A.
106
,
353
360
(
2009
).
18.
J.
Cai
,
G. G.
Guerreschi
, and
H. J.
Briegel
, “
Quantum control and entanglement in a chemical compass
,”
Phys. Rev. Lett.
104
,
220502
(
2010
).
19.
E. M.
Gauger
,
E.
Rieper
,
J. J.
Morton
,
S. C.
Benjamin
, and
V.
Vedral
, “
Sustained quantum coherence and entanglement in the avian compass
,”
Phys. Rev. Lett.
106
,
040503
(
2011
).
20.
J. N.
Bandyopadhyay
,
T.
Paterek
, and
D.
Kaszlikowski
, “
Quantum coherence and sensitivity of avian magnetoreception
,”
Phys. Rev. Lett.
109
,
110502
(
2012
).
21.
H. J.
Hogben
,
T.
Biskup
, and
P. J.
Hore
, “
Entanglement and sources of magnetic anisotropy in radical pair-based avian magnetoreceptors
,”
Phys. Rev. Lett.
109
,
220501
(
2012
).
22.
C. Y.
Cai
,
Q.
Ai
,
H. T.
Quan
, and
C. P.
Sun
, “
Sensitive chemical compass assisted by quantum criticality
,”
Phys. Rev. A
85
,
022315
(
2012
).
23.
H. G.
Hiscock
,
S.
Worster
,
D. R.
Kattnig
,
C.
Steers
,
Y.
Jin
,
D. E.
Manolopoulos
,
H.
Mouritsen
, and
P. J.
Hore
, “
The quantum needle of the avian magnetic compass
,”
Proc. Natl. Acad. Sci. U. S. A.
113
,
4634
4639
(
2016
).
24.
J.
Xu
,
L. E.
Jarocha
,
T.
Zollitsch
,
M.
Konowalczyk
,
K. B.
Henbest
,
S.
Richert
,
M. J.
Golesworthy
,
J.
Schmidt
,
V.
Déjean
,
D. J. C.
Sowood
,
M.
Bassetto
,
J.
Luo
,
J. R.
Walton
,
J.
Fleming
,
Y.
Wei
,
T. L.
Pitcher
,
G.
Moise
,
M.
Herrmann
,
H.
Yin
,
H.
Wu
,
R.
Bartölke
,
S. J.
Käsehagen
,
S.
Horst
,
G.
Dautaj
,
P. D. F.
Murton
,
A. S.
Gehrckens
,
Y.
Chelliah
,
J. S.
Takahashi
,
K.-W.
Koch
,
S.
Weber
,
I. A.
Solov’yov
,
C.
Xie
,
S. R.
Mackenzie
,
C. R.
Timmel
,
H.
Mouritsen
, and
P. J.
Hore
, “
Magnetic sensitivity of cryptochrome 4 from a migratory songbird
,”
Nature
594
,
535
540
(
2021
).
25.
G.
Fleissner
,
E.
Holtkamp-Rötzler
,
M.
Hanzlik
,
M.
Winklhofer
,
G.
Fleissner
,
N.
Petersen
, and
W.
Wiltschko
, “
Ultrastructural analysis of a putative magnetoreceptor in the beak of homing pigeons
,”
J. Comp. Neurol.
458
,
350
360
(
2003
).
26.
C. V.
Mora
,
M.
Davison
,
J.
Martin Wild
, and
M. M.
Walker
, “
Magnetoreception and its trigeminal mediation in the homing pigeon
,”
Nature
432
,
508
511
(
2004
).
27.
D.
Heyers
,
M.
Zapka
,
M.
Hoffmeister
,
J. M.
Wild
, and
H.
Mouritsen
, “
Magnetic field changes activate the trigeminal brainstem complex in a migratory bird
,”
Proc. Natl. Acad. Sci. U. S. A.
107
,
9394
9399
(
2010
).
28.
C. D.
Treiber
,
M. C.
Salzer
,
J.
Riegler
,
N.
Edelman
,
C.
Sugar
,
M.
Breuss
,
P.
Pichler
,
H.
Cadiou
,
M.
Saunders
,
M.
Lythgoe
,
J.
Shaw
, and
D. A.
Keays
, “
Clusters of iron-rich cells in the upper beak of pigeons are macrophages not magneto-sensitive neurons
,”
Nature
484
,
367
370
(
2012
).
29.
S.
Qin
,
H.
Yin
,
C.
Yang
,
Y.
Dou
,
Z.
Liu
,
P.
Zhang
,
H.
Yu
,
Y.
Huang
,
J.
Feng
,
J.
Hao
,
J.
Hao
,
L.
Deng
,
X.
Yan
,
X.
Dong
,
Z.
Zhao
,
T.
Jiang
,
H.-W.
Wang
,
S.-J.
Luo
, and
C.
Xie
, “
A magnetic protein biocompass
,”
Nat. Mater.
15
,
217
226
(
2016
).
30.
X.
Zhao
,
H.-B.
Chen
,
L.-H.
Lu
, and
Y.-Q.
Li
, “
A mechanism of compass-free migratory navigation
,”
J. Phys. D: Appl. Phys.
55
,
245004
(
2022
).
31.
J.
Lenz
and
S.
Edelstein
, “
Magnetic sensors and their applications
,”
IEEE Sens. J.
6
,
631
649
(
2006
).
32.
C. D.
Tesche
and
J.
Clarke
, “
dc SQUID: Noise and optimization
,”
J. Low Temp. Phys.
29
,
301
331
(
1977
).
33.
R. T.
Wakai
and
D. J.
Van Harlingen
, “
Signal and white noise properties of edge junction dc SQUID’s
,”
Appl. Phys. Lett.
52
,
1182
(
1988
).
34.
T.
Ryhänen
,
H.
Seppä
,
R.
Ilmoniemi
, and
J.
Knuutila
, “
SQUID magnetometers for low-frequency applications
,”
J. Low Temp. Phys.
76
,
287
386
(
1989
).
35.
M.
Hirota
,
T.
Furuse
,
K.
Ebana
,
H.
Kubo
,
K.
Tsushima
,
T.
Inaba
,
A.
Shima
,
M.
Fujinuma
, and
N.
Tojyo
, “
Magnetic detection of a surface ship by an airborne LTS SQUID MAD
,”
IEEE Trans. Appl. Supercond.
11
,
884
887
(
2001
).
36.
S.
Xu
,
C. W.
Crawford
,
S.
Rochester
,
V.
Yashchuk
,
D.
Budker
, and
A.
Pines
, “
Submillimeter-resolution magnetic resonance imaging at the Earth’s magnetic field with an atomic magnetometer
,”
Phys. Rev. A
78
,
013404
(
2008
).
37.
T. R.
Clem
,
C. P.
Foley
, and
M. N.
Keene
, “
SQUIDs for geophysical survey and magnetic anomaly detection
” in
The SQUID Handbook: Applications of SQUIDs and SQUID Systems
, edited by
J.
Clarke
and
A. I.
Braginski
(
Wiley VCH
,
2006
), Vol. II, pp.
481
543
.
38.
D.
Budker
and
M.
Romalis
, “
Optical magnetometry
,”
Nat. Phys.
3
,
227
234
(
2007
).
39.
H. B.
Dang
,
A. C.
Maloof
, and
M. V.
Romalis
, “
Ultrahigh sensitivity magnetic field and magnetization measurements with an atomic magnetometer
,”
Appl. Phys. Lett.
97
,
151110
(
2010
).
40.
W. C.
Griffith
,
S.
Knappe
, and
J.
Kitching
, “
Femtotesla atomic magnetometry in a microfabricated vapor cell
,”
Opt. Express
18
,
27167
27172
(
2010
).
41.
A.
Ben-Kish
and
M. V.
Romalis
, “
Dead-zone-free atomic magnetometry with simultaneous excitation of orientation and alignment resonances
,”
Phys. Rev. Lett.
105
,
193601
(
2010
).
42.
G.
Oelsner
,
V.
Schultze
,
R.
IJsselsteijn
,
F.
Wittkämper
, and
R.
Stolz
, “
Sources of heading errors in optically pumped magnetometers operated in the Earth’s magnetic field
,”
Phys. Rev. A
99
,
013420
(
2019
).
43.
G.
Bao
,
A.
Wickenbrock
,
S.
Rochester
,
W.
Zhang
, and
D.
Budker
, “
Suppression of the nonlinear Zeeman effect and heading error in Earth-field-range alkali-vapor magnetometers
,”
Phys. Rev. Lett.
120
,
033202
(
2018
).
44.
H.
Katsura
,
N.
Nagaosa
, and
A. V.
Balatsky
, “
Spin current and magnetoelectric effect in noncollinear magnets
,”
Phys. Rev. Lett.
95
,
057205
(
2005
).
45.
M.
Mochizuki
, “
Spin-wave modes and their intense excitations effects in skyrmion crystals
,”
Phys. Rev. Lett.
108
,
017601
(
2012
).
46.
M. O.
Scully
and
M. S.
Zubairy
,
Quantum Optics
(
Cambridge University Press
,
Cambridge
,
1997
).
47.
S.-Y.
Zhu
and
M. O.
Scully
, “
Spectral line elimination and spontaneous emission cancellation via quantum interference
,”
Phys. Rev. Lett.
76
,
388
391
(
1996
).
48.
Z.
Guo
,
S.
Xu
,
X.
Chen
,
C.
Wang
,
P.
Yang
,
S.
Qin
,
C.
Zhao
,
F.
Fei
,
X.
Zhao
,
P.-H.
Tan
,
J.
Wang
, and
C.
Xie
, “
Modulation of MagR magnetic properties via iron–sulfur cluster binding
,”
Sci. Rep.
11
,
23941
(
2021
).
49.
A. S.
Parmagnani
,
S.
D’Alessandro
, and
M. E.
Maffei
, “
Iron-sulfur complex assembly: Potential players of magnetic induction in plants
,”
Plant Sci.
325
,
111483
(
2022
).
50.
Y.
Zhou
,
T.
Tong
,
M.
Wei
,
T.
Tong
,
M.
Wei
,
P.
Zhang
,
F.
Fei
,
X.
Zhou
,
Z.
Guo
,
J.
Zhang
,
H.
Xu
,
L.
Zhang
,
S.
Wang
,
J.
Wang
,
T.
Cai
,
X.
Zhang
, and
C.
Xie
, “
Towards magnetism in pigeon MagR: Iron- and iron-sulfur binding work indispensably and synergistically
,”
Zool. Res.
44
,
142
152
(
2023
).
51.
H.-B.
Chen
,
Y.
Zhou
, and
Y.-Q.
Li
, “
Bond distortion effects and electric orders in spiral multiferroic magnets
,”
J. Phys.: Condens. Matter
25
,
286004
(
2013
).