zotero/storage/M6ECNMT3/.zotero-ft-cache

647 lines
82 KiB
Plaintext
Raw Blame History

This file contains invisible Unicode characters

This file contains invisible Unicode characters that are indistinguishable to humans but may be processed differently by a computer. If you think that this is intentional, you can safely ignore this warning. Use the Escape button to reveal them.

This file contains Unicode characters that might be confused with other characters. If you think that this is intentional, you can safely ignore this warning. Use the Escape button to reveal them.

Skip to content
De Gruyter
EN
0
Subjects
For Authors
Services
Publications
About
BY 4.0 license Open Access Published by De Gruyter Open Access September 16, 2023
The reciprocal linear effect, a new optical effect of the Sagnac type
Gianfranco Spavieri and Espen Gaarder Haug EMAIL logo
From the journal Open Physics
https://doi.org/10.1515/phys-2023-0110
Cite this Share this
1
1 citation on Dimensions.
Abstract
The Sagnac effect can be demonstrated with light propagating either along a circular contour or, as done by Wang et al ., back and forth along a linear contour. In the linear Sagnac effect, the emitterreceiver device is in motion relative to the contour where light propagates. In the reciprocal linear Sagnac effect (RLSE), the device is stationary and the contour is in motion. When the contour changes direction of motion, some special features of the linear Sagnac effect are not fully reciprocal to the RLSE, which foresees variations of the first order in v c in the round-trip time taken by a light signal to cover the contour. The RLSE can be tested with present technology and, if confirmed experimentally, it might have interesting technological applications. Presently, it can be important for testing light-speed invariance, simultaneity, and the relativity principle.
Keywords: light propagation ; Sagnac effect ; one-way speed of light ; relative simultaneity ; Lorentz invariance
1 Introduction
The experiment related to the circular Sagnac effect [ 1 ], shown in Figure 1 (a), was performed in 1913, and the one corresponding to the linear effect, shown in Figure 1 (b), was accomplished by Wang et al . [ 2 ] in 2003. In both effects, the measuring device C * (emitterreceiver clock or interferometer) is in motion relative to a stationary contour, and C * measures the difference Δ T of the round-trip times of two light signals counterpropagating around the contour. Following Post [ 3 ], Δ T is given by,
(1) Δ T = T ⇐ T ⇒ = 2 L γ ( c v ) 2 L γ ( c + v ) = 4 γ v L c 2 = 2 γ v P c 2 ,
where T ⇐ and T ⇒ represent the round-trip time of the co- and counter-moving light signals (or photons) along the contour of perimeter P = 2 π r or 2 L in the circular and linear effects, respectively. For the circular Sagnac effect, with v = ω r , result ( 1 ) is usually expressed [ 3 ] as Δ T = 4 ω ⋅ A c 2 , where A is the area enclosed by the light path.
Figure 1 (a) In the circular Sagnac effect, two counter-propagating photons are emitted from the device C * {{\rm{C}}}^{* } and travel along the circumference of the rotating platform (only a single photon is shown). C * {{\rm{C}}}^{* } measures the difference Δ T \Delta T of the arrival times after a round trip. (b) In the linear Sagnac effect, the counter-propagating photons travel in an optical fiber that may slide frictionless around the two pulleys A and B. The segment AC* of length D D represents the initial position of device C * {{\rm{C}}}^{* } relative to A, the left end point of the contour. If AC* = = D > 2 v L c = ( v c ) P D\gt 2vL/c=\left(v/c)P , the counter-moving photon performs a round trip and gets back to C * {{\rm{C}}}^{* } when this is still moving on the contour lower section.
Figure 1
(a) In the circular Sagnac effect, two counter-propagating photons are emitted from the device C * and travel along the circumference of the rotating platform (only a single photon is shown). C * measures the difference Δ T of the arrival times after a round trip. (b) In the linear Sagnac effect, the counter-propagating photons travel in an optical fiber that may slide frictionless around the two pulleys A and B. The segment AC* of length D represents the initial position of device C * relative to A, the left end point of the contour. If AC* = D > 2 v L c = ( v c ) P , the counter-moving photon performs a round trip and gets back to C * when this is still moving on the contour lower section.
In the linear Sagnac effect of Figure 1 (b), the arm AB of the contour is stationary and the measuring device C * is moving clockwise with uniform speed v along the contour, going from the lower to the upper section of the contour and vice versa. While sliding around the pulley of radius R during the short time η , the device C * changes the direction of motion at the pulley A (or B). Locally, the speed v of C * relative to the contour is always constant. In the linear experiment performed by Wang et al . [ 2 ], the device C * is always in uniform rectilinear motion on the lower contour section during the round-trip time T ≃ T ⇐ ≃ T ⇒ . Therefore, in this experiment, C * does not turn around the pulleys A or B. However, since the relative speed v between C * and the contour is constant, there is no reason to suppose that the result differs from the theoretical prediction ( 1 ) if C * turns around the pulley during the round-trip time T .
In Section 2, we consider the reciprocal linear Sagnac effect (RLSE), shown in Figure 2 , where the device C * is stationary and the contour is in relative motion. We find that, for counter-propagating light signals, the same observable Δ T in ( 1 ) is foreseen in the RLSE and, thus, with regard to Δ T , the RLSE is reciprocal to the linear Sagnac effect.
Figure 2 In the RLSE, the emitterreceiver C * {{\rm{C}}}^{* } is stationary and the contour of length ≃ 2 L = P \simeq 2L=P is moving with a relative speed v v . (a) The device C * {{\rm{C}}}^{* } emits a photon that travels on the contour lower section from the position AC * = D < 2 v L c 2 {\rm{AC}}* =D\lt 2vL/{c}^{2} . (b) When point A reaches C * {{\rm{C}}}^{* } , the contour changes direction of motion while sliding in the perpendicular direction at speed v v on C * {{\rm{C}}}^{* } for a short negligible time interval η \eta . (c) After the contour has resumed its motion in the horizontal direction, after a round-trip, C * {{\rm{C}}}^{* } receives the returning photon on the contour upper section.
Figure 2
In the RLSE, the emitterreceiver C * is stationary and the contour of length ≃ 2 L = P is moving with a relative speed v . (a) The device C * emits a photon that travels on the contour lower section from the position AC * = D < 2 v L c 2 . (b) When point A reaches C * , the contour changes direction of motion while sliding in the perpendicular direction at speed v on C * for a short negligible time interval η . (c) After the contour has resumed its motion in the horizontal direction, after a round-trip, C * receives the returning photon on the contour upper section.
Nevertheless, within the context of standard special relativity based on light speed invariance, there are observable features of the linear Sagnac effect that are not reciprocal to those of the RLSE. In fact, when in the interval T ⇒ (or T ⇐ ) the device C * turns around the pulley in the linear Sagnac effect of Figure 1 (b), for the corresponding RLSE of Figure 2 , the Lorentz transformations (LTs) foresee variations in the first order in v c for the interval T ⇒ (or T ⇐ ), which differ from that given by Eq. ( 1 ). Thus, if experimentally confirmed, the RLSE represents a new optical effect, which differs from the standard Sagnac effect.
In Section 3, we describe a realistic experiment, feasible with present technology, which can measure the mentioned variations related to the velocity v .
There are no problems in interpreting the circular and linear Sagnac effect in the inertial frame of reference S c of the stationary contour where the speed of light is assumed to be c , while different interpretations of the Sagnac effects are given [ 4 12 ] in the frame S C * comoving with the device C * .
Some of the calculations that support our results are presented in the Appendix. Moreover, in the Appendix, we discuss some of the interpretations of the Sagnac effects in the wider scenario of relativistic theories [ 4 13 ], involving tests on simultaneity confirming that relative and absolute simultaneity are not physically equivalent. In this scenario, we show how the RLSE can be used as a test of light speed invariance.
2 The reciprocal linear Sagnac effect
In the RLSE of Figure 2 , the measuring device C * (clock or interferometer) is stationary, while the whole contour is moving back and forth with uniform speed v relative to C * . Ideally, the motion of the contour should be such that, locally, the relative speed between the contour and C * would always be v . In Figure 2 (a), the contour is moving with uniform motion relative to C * , which stays on the contour lower section until pulley A (not shown) reaches C * . At this moment, since C * is stationary, the contour has to move slightly in the direction perpendicular to AB during the short time η when sliding around C * at A, as shown in Figure 2 (b). After changing direction of motion, the contour is moving relative to C * , now on the upper section as shown in Figure 2 (c).
If the pulleys have diameter 2 R , the motion in the perpendicular direction 2 H ≃ 2 R takes place during the finite time interval η ≃ π R v when the contour is accelerating and changing its direction of motion relative to C * . However, for the purpose of simplifying calculations, we omit the process taking place during the negligible short finite interval η by assuming that the round-trip time T is much greater than η ( T ≫ η and L ≫ η v ≃ 2 H ≃ 2 R , while ( v c ) L ≫ η c ). Thus, in the following, we consider the simple ideal case of a linear contour moving back and forth in the direction AB at the uniform speed v relative to C * . However, in the Appendix, we consider also the more realistic case of a rectangular contour where η is not negligible and the motion perpendicular to AB is taken into account.
Denoting by P = 2 L the perimeter of the contour, assuming the light speed to be c and using the LT, we find the following results valid for the RLSE:
(a) AC * = D > 2 v L c = ( v c ) P . C * remains always on one of the sections of the contour in the interval T .
The segment AC * of length D represents the initial position of device C * relative to A, the left end point of the contour. For light counter-propagation (photon moving counter-clockwise) starting from the device C * , when D > 2 v L c 2 the description of the RLSE is reciprocal to that shown in ( Figure 1 (b)), where C * remains always on the lower section of the contour, while the photon performs a round trip in the interval T ⇒ and gets back to C * . The Lorentz-contracted moving contour has length L γ and the round-trip times for counter-, T ⇒ , and co-propagation, T ⇐ , have the expected values,
(2) T ⇒ = 2 L γ ( c + v ) = 2 γ L c ( 1 v c ) T ⇐ = 2 L γ ( c v ) = 2 γ L c ( 1 + v c ) Δ T = T ⇐ T ⇒ = 4 γ v L c 2 = 4 γ v L c 2 = 2 γ v P c 2 ,
where in Eq. ( 2 ), T ⇐ , T ⇒ , and Δ T are the same as given in Eq. ( 1 ). In some cases, we may approximate the results to the first order in v c to simplify calculations. In Appendix (A.3), we calculate the round-trip interval T ⇒ for the counter-propagating photon emitted by C * from the distance AC * = D > 2 v L c , and show in ( A15 ) that T ⇒ is independent of D when C * stays on the same track (lower contour section, in this case) during the interval T ⇒ .
In this case ( D > 2 v L c ), the RLSE foresees results that are the same as those of the standard linear Sagnac effect. Moreover, the symmetry is such that T ⇐ is the same as T ⇒ by changing v to v . It follows that, for AC * = D > 2 v L c , results ( 2 ) for the RLSE are equivalent to results ( 1 ) of the standard Sagnac effect. In fact, in agreement with the principle of relativity, provided that C * and the contour are in uniform relative motion in the interval T ⇒ , the two effects are the same when observed either from the rest frame of the contour or the device C * .
(b) AC * = D < 2 v L c . C * and the contour change their direction of relative motion in the interval T .
In this case, for light counter-propagation, the successive positions of the contour relative to C * are shown in Figure 2 , indicating that C * and the contour change their direction of relative motion during the round-trip time.
Special case D = D 0 = v L c . Starting from D 0 γ (where the factor γ takes into account the length contraction of the moving contour) on the lower section of the contour, the photon travels at speed c toward the moving point B that, at time t , is at the distance L γ D 0 γ + v t = ( L γ ) ( 1 v c ) + v t . Then, the signal reaches point B when c t = ( L γ ) ( 1 v c ) + v t , i.e. , after the time interval t out = L ( γ c ) . Moreover, since point A is moving toward C * at speed v from the initial distance AC * = D 0 , point A reaches C * when v t = D 0 γ = v L ( γ c ) , i.e. , at t out = L ( γ c ) . Hence, the two events “photon at B” and “A at the position of C * ” are simultaneous in the reference frame S C * where C * is at rest, as shown in Figure 2 (b). The return time interval of the photon from B to C * on the contour upper section is obviously t ret = L ( γ c ) , the return interval t ret being independent of the motion of the contour. Then, in the RLSE of Figure 2 (b), the return interval t ret is the same regardless of whether the contour keeps moving to the right or starts moving to the left of the stationary C * .
Working out the calculations also for the co-propagating signal, for T ⇐ , we find the exact result ( A18 ) of Appendix (A.3). Then,
(3) T ⇐ = 2 L γ c + 4 γ v L c 2 , T ⇒ = 2 L γ c Δ T RLSE = T ⇐ T ⇒ = 4 γ v L c 2 = 2 γ v P c 2 = Δ T = invariant .
We see that, in the special case considered earlier where C * is initially on the lower and then on the upper section of the moving contour, the round-trip time T ⇒ in Eq. ( 3 ) is greater than the corresponding value in Eq. ( 2 ). However, also the round-trip time T ⇐ turns out to be greater than the corresponding value in ( 2 ). Then, for counter-propagating light signals, the difference Δ T RLSE in Eq. ( 3 ) is still the same as in Eq. ( 2 ) and, as far as Δ T is concerned, Δ T = Δ T RLSE = invariant and the RLSE is equivalent to the standard linear Sagnac effect.
Special features of the RLSE. The important interesting feature of the RLSE is that in Eq. ( 3 ) the round-trip time T ⇒ of a single photon ( i.e. , the counter-propagating one) differs from the corresponding T ⇒ in Eq. ( 2 ). The difference persists as long as 0 < D < 2 v L c and C * moves from the lower to the upper section during the round-trip time T ⇒ . Then, the function T ⇒ = T ⇒ ( D ) varies from the minimum value given in Eq. ( 2 ) (for D = 0 or D > 2 v L c ) to the maximum one given in Eq. ( 3 ) (for D = D 0 = v L c ). We must stress that T ⇒ is an observable measurable by clock C * and, for the standard linear Sagnac effect, the theory foresees T ⇒ = 2 L γ ( c + v ) , independent of the value of D . Analogous considerations can be made for T ⇐ .
Thus, results ( 2 ) and ( 3 ) indicate that there are special features of the RLSE not equivalent to those of the standard Sagnac effect. In case (b), when velocity variations are involved and C * and the contour change their direction of relative motion in the interval T , the differences indicate that the relativity principle does not hold. The differences between the two effects are interpreted below in the context of special relativity and the LT, showing that can be linked to relative simultaneity.
Since the standard RLSE, where Δ T = T ⇐ T ⇒ is measured, is equivalent to the standard linear Sagnac effect, the usual approach based on measuring Δ T = Δ T RLSE = invariant does not reveal the special features of the RLSE. Then, for the purpose of pointing out these special features foreseen by the LT, we consider below the special reciprocal linear Sagnac effect (S-RLSE), and show how the S-RLSE can be tested and represents an optical effect with properties that differ from those of the standard linear Sagnac effect.
Testing the S-RLSE with an experiment
We pointed out that the round-trip T ⇒ of the counter-propagating signal is an observable that can be measured by the single clock C * . Therefore, the special features of the RLSE can be revealed with this type of measurements. In general, for precise measurements, we need to compare the observable T ⇒ with the corresponding observable T * of a signal on a different light path. In the case of the S-RLSE, for the optical light path of the co-propagating photon, we choose a contour with the same perimeter P = 2 L , but stationary in the device frame S C * .
Hence, the main difference between the RLSE and the S-RLSE is that in the RLSE, we have T * = T ⇐ , while in the S-RLSE, the round-trip time interval of the co-propagating signal is now T * = 2 L c . With our choice, by means of ( 1 ), for the standard Sagnac effect, we have Δ T * Sagnac = T * T ⇒ = 2 L c 2 γ ( L c ) ( 1 + v c ) ≃ Δ Sagn = 2 v L c 2 constant and independent of D . Instead, for the S-RLSE, we have Δ T * = Δ T * ( D ) = T * T ⇒ = T * T ⇒ ( D ) dependent on D .
With the help of ( 2 ) and ( 3 ), we find,
(4) D ≳ 2 v L c ⇒ Δ T * ( D ) = T * T ⇒ = 2 L c 2 γ L ( 1 v c ) c ≃ 2 v L c 2 = Δ T * Sagnac = Δ Sagn D = D 0 ≃ v L c ≲ 2 v L c ⇒ Δ T * ( D 0 ) = T * T ⇒ ( D 0 ) ≃ 2 L c 2 L c = 0 .
According to results ( 4 ), the function Δ T * ( D ) is constant for D ≳ 2 v L c and given by 2 v L c 2 = Δ Sagn = Δ T * Sagnac , which is the same for the RLSE and the linear Sagnac effect. However, for D ≲ 2 v L c , the function Δ T * ( D ) varies from the maximum values 2 v L c 2 ( D = 0 and D = 2 v L c ) to zero ( D = D 0 = v L c ), as shown in Figure 4 . Analogous results, given in the following section and in the Appendix, are obtained when light propagates in a medium of refractive index n .
Thus, by assuming light speed invariance and the LT, the S-RLSE foresees the variant Δ T * ( D ) of ( 4 ) that can be observed. The variant special feature of ( 4 ) is discussed in the Appendix within the wider scenario of relativistic theories. To fully understand the difference between the D -dependent S-RLSE results ( 4 ) and the D -independent results of the linear Sagnac effect in Eq. ( 2 ) and in Eq. ( 1 ), we consider the following corresponding interpretations.
2.1 Interpreting the RLSE and the linear Sagnac effect using the LTs
In the linear Sagnac effect and in the RLSE, the measurements are made by device C * . Why are the round-trip time intervals of the propagating signals different in the two effects? Let us, then, consider an observer comoving with clock C * (let us call it “observer C * ” or simply “ C * ”) and check first what happens in the linear Sagnac effect of Figure 1 (b) when C * is near point A and, during the round-trip interval T ⇒ , changes the direction of motion in the negligible time interval η moving from the lower to the upper section of the contour. While on the lower section, C * is comoving with the inertial frame S ″ and with the inertial frame S when on the upper section.
If the counter-propagating photon is emitted by C * at AC * = D 0 = ( v c ) L when in the lower section frame S ″ , C * reaches A when, simultaneously, the photon reaches point B, as shown in Figure 3 . Then, for the relative position of photon and C * , the situation is the same as that in the RLSE of Figure 2 (b).
Figure 3 Effect of relative simultaneity observed by C * {{\rm{C}}}^{* } in the linear Sagnac effect. The two events “ C * {{\rm{C}}}^{* } at A” and “photon at B” are simultaneous on frame S ″ {S}^{^{\prime\prime} } comoving with C * {{\rm{C}}}^{* } when on the contour lower section. After turning around the pulley in the negligible time interval η \eta and changing velocity, C * {{\rm{C}}}^{* } starts comoving with frame S {S}^{^{\prime} } on the upper section at t = t ″ = 0 {t}^{^{\prime} }={t}^{^{\prime\prime} }=0 . However, due to relative simultaneity between S {S}^{^{\prime} } and S ″ {S}^{^{\prime\prime} } , for observer C * {{\rm{C}}}^{* } on S {S}^{^{\prime} } , the photon is no longer at B but is at K already.
Figure 3
Effect of relative simultaneity observed by C * in the linear Sagnac effect. The two events “ C * at A” and “photon at B” are simultaneous on frame S ″ comoving with C * when on the contour lower section. After turning around the pulley in the negligible time interval η and changing velocity, C * starts comoving with frame S on the upper section at t = t ″ = 0 . However, due to relative simultaneity between S and S ″ , for observer C * on S , the photon is no longer at B but is at K already.
However, when (after reaching A and turning around the pulley) C * starts comoving with frame S on the upper section, we can see in Figure 3 that now the position of the photon is not at B, but at K, being AK ≃ L ( 1 2 v c ) < L and, thus, the situation differs from that of Figure 2 (b) of the RLSE.
The difference is due to the fact that, in the linear Sagnac effect, the mechanism of relative simultaneity takes place when C * changes velocity by moving from frame S ″ on the lower section to the frame S on the upper section [ 7 ]. The two events, “ C * at A” and “photon at B,” are simultaneous for observer C * on frame S ″ in the lower section, but no longer simultaneous for C * on frame S in the upper section because S is in motion with speed ≃ 2 v relative to S ″ .
As shown in Figure 3 , in the linear Sagnac effect, clock C * is reached by the photon returning on the upper section after the proper time interval τ ret ≃ ( L c ) ( 1 2 v c ) and the round-trip time interval is, as expected, ( T ⇒ ) Sagnac = τ out + τ ret ≃ L c + ( L c ) ( 1 2 v c ) = 2 L ( c + v ) . In this case, relative to C * , in the return trip, the photon covers at speed c the shorter distance L ( 1 2 v c ) only.
However, in the case of the RLSE ( Figure 2 (b)), C * is stationary and, since the mechanism of relative simultaneity does not apply, in the return trip, the photon covers at speed c the full distance ≃ L in the time interval t ret = L ( γ c ) ≃ L c . Thus, for the special case D = D 0 = v L c 2 , in the RLSE, the round-trip time is ( T ⇒ ) RLSE ≃ 2 L c , which is greater than the corresponding one ( T ⇒ ) Sagnac ≃ 2 L ( c + v ) of the linear Sagnac effect.
As shown in ref. [ 7 ], in the case of the linear Sagnac effect, the “time gap” δ t ≃ 2 v L c 2 , due to relative simultaneity between S and S ″ , corresponds to the length difference c δ t ≃ 2 v L c not covered by the photon when C * is on the upper section. Therefore, we find that ( T ⇒ ) RLSE ≃ ( T ⇒ ) Sagnac + δ t , showing that the difference can be related to relative simultaneity.
In general, in the linear Sagnac effect, the round-trip time T of light signals is independent of D and the stationary contour shape, while in the RLSE, the round-trip time T depends on D and on the shape of the moving contour. Then, the RLSE represents an optical effect not fully equivalent to the linear Sagnac effect.
3 Experiment testing the S-RLSE
Here, we discuss the results obtained for the rectangular contour of Figure 4 (shown also in Figure A1 of the Appendix) of sizes L and H . With H ≃ 2 R , where R is the radius of the pulley in the linear Sagnac effect, the shape of this contour represents more realistically the RLSE.
Figure 4 Time difference Δ T * ( D ) \Delta {T}^{* }\left(D) for the counter-moving photon in the S-RLSE. For the Sagnac effect, Δ T \Delta T * = v P c 2 =vP/{c}^{2} is constant and represented by the dotted line. For the S-RLSE, starting from right to left with C * {{\rm{C}}}^{* } on the contour lower section L L from the position D > 2 v L c 2 D\gt 2vL/{c}^{2} , we have Δ T * = v P c 2 \Delta {T}^{* }=vP/{c}^{2} until D ≃ 2 v L c D\simeq 2vL/c . When D < 2 v L c D\lt 2vL/c and until D = D 0 = v L c D={D}_{0}=vL/c , Δ T * ( D ) \Delta {T}^{* }\left(D) decreases up to v L c 2 vL/{c}^{2} . Then, Δ T \Delta T * increases and reaches again the value v P c 2 vP/{c}^{2} at D = 0 D=0 and afterwards when C * {{\rm{C}}}^{* } is on the left side H H of the moving contour. The broken line refers to the S-RLSE of Figure 2, where P = 2 L P=2L and the dip is 2 v L c 2 2vL/{c}^{2} with Δ T * = 0 \Delta {T}^{* }=0 at D = D 0 D={D}_{0} .
Figure 4
Time difference Δ T * ( D ) for the counter-moving photon in the S-RLSE. For the Sagnac effect, Δ T * = v P c 2 is constant and represented by the dotted line. For the S-RLSE, starting from right to left with C * on the contour lower section L from the position D > 2 v L c 2 , we have Δ T * = v P c 2 until D ≃ 2 v L c . When D < 2 v L c and until D = D 0 = v L c , Δ T * ( D ) decreases up to v L c 2 . Then, Δ T * increases and reaches again the value v P c 2 at D = 0 and afterwards when C * is on the left side H of the moving contour. The broken line refers to the S-RLSE of Figure 2 , where P = 2 L and the dip is 2 v L c 2 with Δ T * = 0 at D = D 0 .
For the rectangular contour of Figure A1 , we assume that the contour slides continuously at the local speed v relative to C * . This is a mechanical difficulty that, in principle, can be surmounted, although it increases the cost of the experiment in comparison to the standard Sagnac experiments. Actually, for the experiment, the rectangular contour can move relative to C * for a short distance, starting with C * on the lower section L and ending on the left side H , as shown in Figure A1 . For the size of the contour to be manageable, it is convenient to have light propagating in an optical fiber with a high refractive index n . In this case, the round-trip time is increased to T ⇒ ≃ n P c .
In the case of light propagation in a medium, we have to distinguish the case when the medium is locally at rest along the contour from the case when the medium is locally at rest with C * and sliding along the contour. The speed of light in the moving medium is obtained from the standard relativistic velocity transformations. The results using the LT, derived in the Appendix (neglecting dispersion), are shown schematically below.
S-RLSE with medium locally at rest with the contour P = 2 L + 2 H
T * = n P c , Δ T * ≃ n 2 v c P c , independent of D
(5) D = D 0 ≃ v n L c 2 , T ⇒ ≃ n P c n 2 v c P c + v L c 2 D = D 0 ≃ v n L c 2 , Δ T * ( D 0 ) = T * T ⇒ ≃ Δ Sagn v L c 2 D > 2 v n L c 2 , T ⇒ ≃ n P c n 2 v c P c D > 2 v n L c 2 , Δ T * ( D ) = T * T ⇒ ≃ Δ Sagn .
S-RLSE with medium locally at rest with C * , P = 2 L + 2 H
T * = n P c , Δ Sagn ≃ v c P c , independent of D
(6) D = D 0 ≃ v n L c 2 , T ⇒ ≃ n P c v c P c + v L c 2 D = D 0 ≃ v n L c 2 , Δ T * ( D 0 ) = T * T ⇒ ≃ Δ Sagn v L c 2 D > 2 v n L c 2 , T ⇒ ≃ n P c v c P c D > 2 v n L c 2 , Δ T * ( D ) = T * T ⇒ ≃ Δ Sagn .
For the S-RLSE with medium locally at rest with device C * , Figure 4 shows Δ T * = Δ T * ( D ) as a function of D . For the Sagnac effect, Δ T * = Δ T * Sagnac = Δ Sagn = v P c 2 is constant. For the S-RLSE, Δ T * ( D ) is represented by the solid line that varies as a function of the initial position D of clock C * . For the S-RLSE, the LT foresee, the observable dip by v L c 2 in the function Δ T * ( D ) . The broken line refers to the S-RLSE of Figure 2 , where P = 2 L and the dip is 2 v L c 2 with Δ T * = 0 at D = D 0 .
Expected precision for the S-RLSE
The variation Δ T * ( D ) may be measured with the techniques of ring interferometry or, after some adaptations, ring lasers [ 3 ]. Ring lasers Sagnac experiments, which can measure the angular velocity ω E of the Earth with a precision δ ω ω E < 1 0 8 , are performed routinely on the Earth [ 14 , 15 ]. For the effect of refractive index of an optical medium on the rotation frequency of ring lasers and resonators, there are various experimental expressions as functions of n , considered by Malykin in ref [ 16 ] (2014).
In relation to the sensitivity of detectors measuring δ τ A and the smallest measurable time interval, there are techniques capable of resolving femtosecond ( 1 0 15 s ) [ 17 ] or even attosecond ( 1 0 18 s ) [ 18 ] pulses of laser light, although better limits may be obtained by means of advanced interferometry.
The dip by v L c 2 , to be measured to confirm the prediction of standard relativity based on the LT, is comparable to that measured in the usual experiments testing the Sagnac effect and, thus, well within the range of the sensitivity of available detectors and not too difficult to observe. In the experiment by Wang et al . [ 2 ], the device C * is in motion relative to the contour. In the S-RLSE, the device C * is stationary and the contour in relative motion. Assuming the same relative velocity, we expect that the sensitivity achievable for our S-RLSE experiment is approximately the same as that achieved in the experiment by Wang et al . [ 2 ]. The main difference with respect to the experiment by Wang et al . is related to the mechanical difficulties involved with the motion of the contour, which implies more an increase in the complexity and cost of the experiment, rather than a loss in sensitivity.
Possible applications of the S-RLSE if confirmed experimentally.
After more than a century, the Sagnac effect is employed in current technology and is used in inertial guidance systems, ring laser gyroscope (extremely sensitive to rotations), and other optical systems. Being not yet experimentally confirmed, it might be premature to indicate what kind of applications the RLSE might have. However, considering that the circular Sagnac effect (sensitive to rotations) can detect very accurately the angular velocity of objects, such as the Earth, the RLSE (sensitive to velocity changes) might be suitable to detect velocity variations and also point out the corresponding direction. The realization of these applications requires, however, to take into account other features related to the RLSE that will be considered in a future contribution.
4 Conclusions
We have described an optical effect, denoted as RLSE, where the emitterreceiver device is stationary and the contour along which light propagates is in motion. The RLSE provides the same results of the standard Sagnac effect for the round-trip time difference Δ T of light signals counter-propagating along the closed contour. However, the RLSE possesses features that are not equivalent to the corresponding ones of the Sagnac effect. When the contour changes velocity, standard special relativity foresees that the round-trip time interval T ⇒ of a counter-moving light signal differs from the corresponding one of the Sagnac effect. The difference is observable with our S-RLSE experiment and given by the dip by v L c 2 in the measured round-trip time difference Δ T * ( D ) = T * T ⇒ of counter-propagating photons, plotted in Figure 4 . The dip may be related to the different role played by relative simultaneity in the linear Sagnac effect and the RLSE.
If confirmed experimentally, the S-RLSE stands to be a new relativistic optical effect. In this case, being sensitive to velocity changes, the S-RLSE might have important technological applications in inertial guidance systems by detecting velocity variations and corresponding direction.
In any event, the different features between the standard linear Sagnac effect and the S-RLSE can be exploited for testing Lorentz and light speed invariance.
Funding information: Our research has been supported by the CDCHTA of the Universidad de Los Andes, Mérida, Venezuela, and the Braingain grant of the International Center for Theoretical Physics (ICTP), Trieste, Italy, for promoting teaching and researching in Venezuela.
Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.
Conflict of interest : The authors state no conflict of interest.
Appendix
A.1 Special reciprocal linear Sagnac effect for light propagation in a medium of refracting index n on a rectangular contour of sides L and H
For the S-RLSE, the round-trip interval T ⇒ differs from the corresponding one of the linear Sagnac effect if, on the contour lower section, to the first order in v c , D ≃ n v L c and the photon reaches the right corner when the left corner reaches C * .
S-RLSE with medium locally at rest with the contour. P = 2 L + 2 H .
For the contour of Figure A1 (a), we calculate with the LT to the first order in v c the time intervals taken by the counter-propagating photon to cover the sides of the rectangular contour in motion with velocity v relative to the stationary C * .
Position 1: Starting from D = v t ≃ n v L c , and using the equation C n t = L D + v t with C n the light speed relative to C * . Then, we have,
C n = c n + v 1 + v n c t 1 = L C n = n L ( 1 + v n c ) c ( 1 + n v c ) ≃ n L c n 2 v L c 2 + v L c 2 .
Position 2: By means of the equation C n t = H v t ,
C n = c n v 1 v n c t 2 = H C n + v ≃ n H ( 1 v n c ) c ≃ n H c v H c 2 .
Position 3: With the equation C n t = L ,
C n ≃ c n t 3 = L C n ≃ n L c .
Position 4: By means of the equation, C n t = H v ( t 2 + t 3 ) ,
C n = c n + v 1 + v n c t 4 = n H ( 1 + v n c ) c ( 1 + n v c ) n 2 v H c 2 n 2 v L c 2 = n H c n 2 v H c 2 + v H c 2 n 2 v H c 2 n 2 v L c 2
T ⇒ = t 1 + t 2 + t 3 + t 4 = n P c n 2 v c P c + v L c 2 .
S-RLSE medium locally at rest with the device C * . P = 2 L + 2 H .
For the contour of Figure A1 (b), we calculate with the LT to the first order in v c the time intervals taken by the counter-propagating photon to cover the optical fiber sliding on the moving rectangular contour.
Position 1: Starting from D = v t , by means of the equation C n t = L D + v t with C n the light speed relative to C * along the fiber at rest, we have,
C n = c n t 1 = L C n = n L c .
Position 2: From the equation C n t = H v t and with the optical fiber at speed 2 v relative to C * ,
C n = c n 2 v 1 2 v n c t 2 = H C n + v ≃ n H ( 1 2 v n c ) c ( 1 n v c ) ≃ n H c + n 2 v H c 2 2 v H c 2 .
Figure A1 (a) Optical fiber of index n n locally at rest on the rectangular contour moving with velocity v v relative to the stationary clock C * {{\rm{C}}}^{* } . The photon starts from C * {{\rm{C}}}^{* } at the distance D D from the contour left side H H and returns to C * {{\rm{C}}}^{* } in the interval T ⇒ {T}_{\Rightarrow } after traveling along the contour sides at different speeds C n {C}_{n} relative to C * {{\rm{C}}}^{* } . (b) Optical fiber of index n n locally at rest with C * {{\rm{C}}}^{* } and sliding on the rectangular contour moving with velocity v v relative to the stationary clock C * {{\rm{C}}}^{* } . In position 1, the fiber in the lower contour section is at rest relative to C * {{\rm{C}}}^{* } . In 2, the sliding fiber has speed 2 v 2v relative to C * {{\rm{C}}}^{* } . In 3, the relative speed is v v and in 4, the fiber is at rest relative to C * {{\rm{C}}}^{* } .
Figure A1
(a) Optical fiber of index n locally at rest on the rectangular contour moving with velocity v relative to the stationary clock C * . The photon starts from C * at the distance D from the contour left side H and returns to C * in the interval T ⇒ after traveling along the contour sides at different speeds C n relative to C * . (b) Optical fiber of index n locally at rest with C * and sliding on the rectangular contour moving with velocity v relative to the stationary clock C * . In position 1, the fiber in the lower contour section is at rest relative to C * . In 2, the sliding fiber has speed 2 v relative to C * . In 3, the relative speed is v and in 4, the fiber is at rest relative to C * .
Position 3: By means of the equation C n t = L and with the optical fiber at speed v ,
C n ≃ c n v 1 v n c t 3 = L C n = n L ( 1 v n c ) c ( 1 n v c ) ≃ n L c + n 2 v L c 2 v L c 2 .
Position 4: From the equation C n t = H v ( t 2 + t 3 ) and with the optical fiber at rest,
C n = c n t 4 = n H c n 2 v H c 2 n 2 v L c 2
T ⇒ = t 1 + t 2 + t 3 + t 4 = P c ( n v c ) + v L c 2 .
A.2 Interpreting optical effects and the RLSE in the scenario of relativistic theories
Several authors [ 4 , 13 , 19 23 ] argue that the LT are physically equivalent to alternative coordinate transformations that differ from the LT by an arbitrary clock synchronization parameter only. Thus, these authors claim that the one-way light speed is undetermined, or conventional, when measured by two spatially separated clocks that can be arbitrarily synchronized. In the framework of relativistic theories [ 13 ], the coordinate transformations alternative to the LT, more often used to describe physical phenomena, are denoted as LTs based on absolute simultaneity (LTA). The LT and the LTA are expressed as,
(A1) L T t = γ ( t v x c 2 ) x = γ ( x v t ) L T A t = t γ x = γ ( x v t ) ,
where in ( A1 ) v indicates the velocity of frame S relative to frame S and the transformations y = y ; z = z are understood. With the factor γ = ( 1 v 2 c 2 ) 1 2 depending on v , LT stands for the LTs, based on standard synchrony and relative simultaneity. LTA stands for the LTs based on absolute synchrony and simultaneity. To the first order in v c , the LTA coincides with the Galileo transformations. The LTA (or ALT in Ref. [ 9 ]) are known in literature also as the TangherliniSelleri transformations [ 6 , 19 , 20 ], used by several authors [ 4 9 ].
Because of the arbitrariness of synchronization and one-way light speed, many physicists consider the LT and LTA to be physically equivalent and, according to them, the LTA can be used to interpret all the experiments supporting standard special relativity [ 13 , 21 23 ]. Certainly, the equivalence should hold in the cases when the one-way light speed is conventional because measured by two spatially separated clocks that are arbitrarily synchronized. However, in the Sagnac effects, the one-way speed of light around the contour can be determined with the single clock C * and no arbitrary clock synchronization is involved. Thus, in relation to the difficulties pointed out in the interpretations of the Sagnac effect [ 6 9 , 11 ], several authors [ 4 9 ] have shown that the difficulties are surmounted if, in lieu of the LT based on relative simultaneity, coordinate transformations based on conservation of simultaneity (LTA) are adopted.
Nevertheless, epistemologists [ 24 ] claim that the basic postulates of a meaningful physical theory must be testable ( i.e. , falsifiable). Then, if one of its basic postulates is not falsifiable, it may be argued by physicists and epistemologists [ 24 ] that the theory is not physically meaningful. If the LT (with relative simultaneity) are equivalent to the LTA (with absolute simultaneity) and the speed of light is conventional, the standard theory of special relativity has a drawback because its fundamental postulate of one-way light speed invariance cannot be tested [ 13 , 19 23 ]. For this reason, in the more recent formulation of special relativity found in mainstream physics journals where light speed is conventional [ 4 ], the constant c in the second postulate is no longer the one-way light speed, but “ the round-trip speed of light ( i.e. , the average speed of light during the round-trip from A to B and then back to A) ,” which is observable.
Nevertheless, standard special relativity is based on the LT and, for determining the good standing of this theory and the LT, it is essential to establish whether the LTA and the LT are, or are not, physically equivalent.
It is worth mentioning that in practically all the experiments supporting special relativity what is being tested are effects related to relativistic time dilation and length contraction, equally foreseen by the LT and the LTA [ 4 9 , 13 ]. Special relativity is a very well tested theory confirmed by high precision experiments [ 13 , 25 28 ]. The experiments performed so far are likely suitable to test [ 13 ] the so-called relativistic effects only, which are of second (or higher) order in v c and are not dedicated experiments capable of testing the special different features of the LT and LTA, relative versus absolute simultaneity, because these differences vanish on average. In literature, there are proposals of dedicated experiments of first order in v c that might test the different features of simultaneity, relative or absolute (see, for example, [ 7 , 20 ]). However, up to now and as far as we know, no experiments of this type have been performed. Yet, as shown below, these different features can be tested by means of the RLSE also.
A.2.1 The LTA and LT foresee different results for the RLSE
In the interpretation of the optical effects of the Sagnac type with the LTA, we are looking for the preferred frame where the one-way light speed can be assumed to be c . The natural choice is given by the contour inertial rest frame S c where Maxwells equations are valid, and the electromagnetic waves, locked on the contour, propagate at speed c . With this choice of the preferred frame, from the contour inertial frame S c , the LTA and LT provide an equivalent interpretation with the same results.
If, for the RLSE of Figure 2 , we choose again the contour to be the preferred frame, then, we may assume that while moving with uniform velocity relative to C * the contour is carrying along the electromagnetic waves, always propagating at the local speed c relative to the contour. With the aforementioned assumptions, we find that in the framework of the LTA, the RLSE is fully reciprocal to the linear Sagnac effect. In fact, save for the negligible time interval η when the contour changes velocity, the observer S c comoving with the contour of Figure 2 is on an inertial frame before and after the interval η . Then, as far as the RLSE is concerned, the result is independent of whether the relative change of motion is performed by C * or the contour. Since simultaneity is conserved with the LTA, when the contour changes velocity, there are no variations in the relative positions of photon and C * , unlike what happens with relative simultaneity in Figure 3 . Therefore, if calculated from the contour frame S c , the invariant proper time interval T ⇒ is the same as in the linear Sagnac effect. However, when calculated from clock frame S C * we have to take into account that with the LTA light speed is no longer invariant.
To verify that, with the LTA, the invariant interval T ⇒ in the RLSE is independent of D and the same as in the linear Sagnac effect, we calculate it from the clock frame S C * for the case when the contour changes velocity in the interval T ⇒ , as in Figure 2 . Let the photon start from C * at D and travel toward point B moving to the right ( Figure 2 (a)). Relative to C * , by addition of velocities, the light speed along the moving contour is ≃ c + v , where for simplicity we use the first-order approximation in v c . In the out trip and at time t , point B is at the position L D + v t to the right of C * and the photon reaches B when ( c + v ) t = L D + v t . Considering first the special case when simultaneity takes place, we set D = v t and, thus, t = t out = L ( c + v ) and D = v t out ≃ v L c . Then, the two events “photon at B” and “A at C * ” are simultaneous in every frame. In the return trip on the upper section ( Figure 2 (b)), the contour is now moving to the left, while the photon travels on the upper section toward the stationary C * at the speed ≃ c + v . With the photon coming from B at L , the return time interval is t ret = L ( c + v ) and the round-trip is T ⇒ = t out + t ret ≃ 2 L ( c + v ) , which is the same as in the linear Sagnac effect shown in ( 2 ).
It can be shown that the same result, independent of D , is obtained in general, e.g. , when the photon reaches B before A reaching C * (or when A reaches C * before the photon reaching B) and when C * is always on the lower (or upper) section of the contour in the round-trip interval T ⇒ .
In conclusion, with the LTA based on absolute simultaneity, the RLSE is fully equivalent to the linear Sagnac effect and the relativity principle is holding. For the S-RLSE experiments described earlier, contrary to the LT, the LTA foresee that the variant (for the LT) function Δ T * ( D ) is (for the LTA) always Δ T * = Δ T * Sagnac , independent of D . It follows that, by means of the S-RLSE, the different predictions of the LT and LTA, and light speed invariance, can be tested.
Another example indicating that the LT and LTA are not equivalent, and cannot be arbitrarily interchanged in the description of physical phenomena, is given by the Thomas precession, discussed below.
A.2.2 Using the Thomas precession to discriminate relative (LT) from absolute (LTA) simultaneity
In the following, we consider the relativistic phenomenon of the Thomas precession [ 29 ] and find that it is predicted by the LT but not by the LTA. Since the spacetime symmetry of the LT differs from that of the LTA, it is not surprising that the two transformations predict different results. Standard special relativity is routinely used in several areas of modern physics employing the symmetry of the LTs based on relative simultaneity and the Thomas precession derived from the LT is well known for predicting the correct spinorbit interaction energy splitting [ 29 , 30 ] and is thus considered to be observable. Moreover, Thomas precession is included in the equation of spin motion of the BMT equation of elementary particle physics [ 31 ]. However, so far there are no direct tests of the Thomas precession.
Deriving the Thomas precession with the LT
Here, we consider the textbook derivation made by Jackson [ 30 ] and indicate schematically the steps that lead to the Thomas precession. In deriving the spin precession for the case of spinorbit interaction, Thomas considers an electron orbiting on a plane at the peripheral velocity v around the nucleus of an atom. Thomas shows that, because of the motion of the electron in its circular orbit with the acceleration a due to the Coulomb field, the electron spin acquires an extra angular velocity ω T of purely relativistic kinematical origin. The Thomas angular velocity ω T , corresponding to the rotation per unit of time of the electron frame, can be calculated from the infinitesimal LT along the electron path (shown in Figure A2 ), indicating that successive transformations consist of a pure boost and a rotation.
Figure A2 An object (electron) is in accelerated motion along a curvilinear path under the action of external forces. The moving frames S {S}^{^{\prime} } and S ″ {S}^{^{\prime\prime} } represent successive rest frames of the object. Starting from the origin of the laboratory frame S S , the velocity of the rest frame S {S}^{^{\prime} } at time t t is v ( t ) = c β {\bf{v}}\left(t)=c\beta and at time t + δ t t+\delta t the velocity of the rest frame S ″ {S}^{^{\prime\prime} } is v ( t + δ t ) = c β + c δ β v\left(t+\delta t)=c\beta +c\delta \beta . The resulting infinitesimal LT from S {S}^{^{\prime} } to S ″ {S}^{^{\prime\prime} } consists of a boost and a rotation, indicating that the object acquires a precession at the Thomas angular velocity ω T {{\boldsymbol{\omega }}}_{T} of purely relativistic kinematical origin independent of other effects.
Figure A2
An object (electron) is in accelerated motion along a curvilinear path under the action of external forces. The moving frames S and S ″ represent successive rest frames of the object. Starting from the origin of the laboratory frame S , the velocity of the rest frame S at time t is v ( t ) = c β and at time t + δ t the velocity of the rest frame S ″ is v ( t + δ t ) = c β + c δ β . The resulting infinitesimal LT from S to S ″ consists of a boost and a rotation, indicating that the object acquires a precession at the Thomas angular velocity ω T of purely relativistic kinematical origin independent of other effects.
To point out the origin of the rotation of the particle rest frame S ″ , in his approach to the Thomas precession, Jackson starts from the LT between the two frames S ″ and S given by (Ref. [ 30 ], Section 11.19),
(A2) x ″ 0 = γ u ( x 0 β ⋅ x ) = γ u ( x 0 β 1 x 1 β 2 x 2 ) x ″ 1 = x 1 + γ u 1 β 2 ( β 1 x 1 + β 2 x 2 ) β 1 γ β 1 x 0 x ″ 2 = x 2 + γ u 1 β 2 ( β 1 x 1 + β 2 x 2 ) β 2 γ β 2 x 0 x ″ 3 = x 3 ,
where, as shown in Figure A2, S ″ is in motion with velocity components u x = u 1 and u y = u 2 relative to the laboratory frame S and in ( A2 ) γ u = ( 1 u 2 c 2 ) 1 2 , u 2 = u 2 1 + u 2 2 , β 1 = u 1 c , and β 2 = u 2 c .
The connection between the two sets of rest frame coordinates, x at time t and x ″ at time t + δ t , corresponding to the frames instantaneously co-moving with the particle in its accelerated motion, are given by,
x = A boost ( β ) x x ″ = A boost ( β + δ β ) x ,
where β = v c and β + δ β = u c = v c + δ β , with the factor γ u = γ + γ 3 β δ β 1 to the first order in δ β . The resulting infinitesimal transformation between S and S ″ is,
(A3) x ″ = A LT x ,
where the transformation matrix,
(A4) A L T = A boost ( β + δ β ) A 1 boost ( β ) ,
is,
(A5) A LT = ( 1 γ 2 δ β 1 γ δ β 2 0 γ 2 δ β 1 1 γ 1 β δ β 2 0 γ δ β 2 γ 1 β δ β 2 1 0 0 0 0 1 ) .
As shown by Jackson [ 30 ], the resulting infinitesimal Lorentz transformation ( A3 ) consists of a boost with velocity c Δ β and a rotation Δ Ω given by,
(A6) Δ Ω = γ 1 β 2 β × δ β .
The rotation is highlighted by the two antisymmetric matrix elements ( A LT ) 23 and ( A LT ) 32 of A LT ,
(A7) A LT = ( ⋯ ⋯ ⋯ ⋯ ⋯ ⋯ γ 1 β δ β 2 ⋯ ⋯ γ 1 β δ β 2 ⋯ ⋯ ⋯ ⋯ ⋯ ⋯ ) ,
indicating the existence of a rotation of the electron rest frame S ″ by Δ Ω about the x 3 axis. Because of the spacetime symmetry of the LT in ( A2 ), the parameter time depends on space, t ″ = t ″ ( t , x ) and t = t ( t , x ) . Therefore, through the matrix multiplication ( A4 ), we can see that the time dependence on space, related to relative simultaneity, is essential for providing the resulting antisymmetric matrix elements ( A LT ) 23 and ( A LT ) 32 of A LT shown in ( A7 ).
A direct consequence of expression ( A6 ) is the resulting Thomas precession acquired by the object at the angular velocity,
(A8) ω T = lim δ t → 0 Δ Ω δ t = γ 2 γ + 1 a × v c 2 .
Then, if the rotating object has a component of a perpendicular to v , the object acquires a Thomas precession of purely kinematical origin, independent of other effects.
The LTA based on absolute simultaneity predict no Thomas precession
The LTs based on absolute simultaneity (LTA) between frame S and S ″ are given by,
(A9) x ″ 0 = γ 1 u x 0 x ″ 1 = x 1 + γ u 1 β 2 ( β 1 x 1 + β 2 x 2 ) β 1 γ β 1 x 0 x ″ 2 = x 2 + γ u 1 β 2 ( β 1 x 1 + β 2 x 2 ) β 2 γ β 2 x 0 x ″ 3 = x 3 ,
which are the same as the LT in ( A2 ), save for the fact that the time parameter does not depend on space (absolute simultaneity) and is given by c τ ″ = x ″ 0 = γ 1 u x 0 , being τ ″ the proper time on the particle rest frame S ″ .
Obviously, the spacetime symmetry of the LT is different from that of the LTA and, therefore, for observables that depend on symmetry, different predictions may be expected from the two different transformations.
Following the procedure adopted earlier, we proceed by calculating the connection,
(A10) x ″ = A LTA x ,
using the LTA (instead of the LT) and derive the corresponding matrix A LTA ( A14 ).
The transformations between frame S and S are,
(A11) x 0 = γ 1 x 0 ⇒ x 0 = γ x 0 x 1 = γ ( x 1 β x 0 ) ⇒ x 1 = γ 1 x 1 + γ β x 0 x 2 = x 2 ⇒ x 2 = x 2 x 3 = x 3 ⇒ x 3 = x 3 .
After substituting x 0 and x 1 given by ( A11 ) in expression ( A9 ), the connection x ″ = A LTA x of ( A10 ) to the first order in δ β is found to be,
(A12) x ″ 0 = γ 1 u x 0 = γ 1 u γ x 0 x ″ 1 = γ u β 1 γ x 0 + ( 1 + γ u 1 β 2 β 1 β 1 ) ( γ 1 x 1 + β γ x 0 ) + γ u 1 β 2 δ β 2 β 1 x 2 x ″ 2 = γ δ β 2 γ x 0 + γ 1 β 2 δ β 2 β 1 ( γ 1 x 1 + β γ x 0 ) + x 2 x ″ 3 = x 3 .
For the matrix elements ( A LTA ) 23 and ( A LTA ) 32 , we may take into account that the velocity of S ″ relative to S is β 1 = β + δ β 1 and β 2 = δ β 2 and γ u ≃ γ to the first order in δ β . Then, ( A12 ) becomes,
(A13) x ″ 0 = γ 1 u γ x 0 x ″ 1 = γ ( β + γ u 1 β 2 β β 1 β 1 γ u β 1 ) x 0 + ( 1 + γ u 1 β 2 β 1 β 1 ) γ 1 x 1 + γ 1 β δ β 2 x 2 x ″ 2 = γ ( γ 1 β δ β 2 β 1 γ δ β 2 ) x 0 + γ 1 γ β δ β 2 x 1 + x 2 x ″ 3 = x 3 .
From the connection x ″ = A LTA x given by Eq. ( A13 ), we find that the elements of the corresponding matrix A LTA relevant to our case are,
(A14) A LTA = ( ⋯ ⋯ ⋯ ⋯ ⋯ ⋯ γ 1 β δ β 2 ⋯ ⋯ γ 1 γ β δ β 2 ⋯ ⋯ ⋯ ⋯ ⋯ ⋯ ) ,
indicating that both the matrix elements ( A LTA ) 23 and ( A LTA ) 32 are positive, not antisymmetric, and do not reflect the existence of a rotation about the x 3 axis, as it is the case for the corresponding elements of the matrix A LT Eq. ( A5 ) derived with the LT. In fact, the matrix elements ( A LTA ) 23 and ( A LTA ) 32 indicate a relative distortion of axes due to length contraction in the direction of motion, but no relative rotation about the x 3 axis.
In conclusion, standard special relativity predicts the existence of the Thomas precession, which is related to the spacetime symmetry properties of the LT based on relative simultaneity. Instead, the LTA, based on conservation of simultaneity, predicts that the Thomas precession does not exist. With the examples of the Thomas precession, the RLSE, and other phenomena [ 6 9 ], we may infer that the LTA are not in general physically equivalent to the LT. Hence, the two transformations and absolute and relative simultaneity cannot be interchanged arbitrarily.
A.3 Calculating with the Lorentz transformations the intervals T ⇐ , T ⇒ and Δ T = T ⇐ T ⇒ for the RLSE
A.3.1 Showing that, for D > 2 v L c and C * on the same (upper or lower) contour section, in the RLSE the round-trip time interval T ⇒ for the counter-moving photon is independent of D and the same as in the standard linear Sagnac effect
With reference to Figures 1 (b) and 2 (a), we assume that the counter-moving photon is emitted by C * when D > 2 v L c , so that C * is always on the lower track during the round-trip interval T ⇒ . In this case, the contour is always moving at speed + v relative to C * . Then, the photon moving at speed c reaches point B when c t = ( L D ) γ + v t , i.e. , at,
t 1 = L D γ ( c v ) = γ ( L D ) ( 1 + v c ) c ,
where we have used the relation 1 γ 2 = ( 1 v c ) ( 1 + v c ) = 1 v 2 c 2 .
Moving now on the upper track, the photon travels from B toward A and reaches A when L γ c t = v t after the interval,
t 2 = L γ ( c + v ) = γ L ( 1 v c ) c .
After reaching A, the photon starts moving on the lower track to return to C*. Since point A has been moving at speed v toward C * during the interval t 1 + t 2 , the position of A relative to C * is now D γ v ( t 1 + t 2 ) and the photon reaches C * when D γ v ( t 1 + t 2 ) = c t , after the interval,
t 3 = D γ c v c ( t 1 + t 2 ) = D γ c v γ c ( ( L D ) ( 1 + v c ) c + L ( 1 v c ) c ) = γ D ( 1 v 2 c 2 ) c v γ c 2 L c + v γ c D ( 1 + v c ) c = γ D ( 1 + v c ) c v γ c 2 L c .
The resulting round-trip interval T ⇒ is,
(A15) T ⇒ = t 1 + t 2 + t 3 = γ ( L D ) ( 1 + v c ) c + γ L ( 1 v c ) c + γ D ( 1 + v c ) c v γ c 2 L c = γ L ( 1 + v c ) c + γ L ( 1 v c ) c v γ c 2 L c γ D ( 1 + v c ) c + γ D ( 1 + v c ) c = 2 γ L ( 1 v c ) c = 2 L γ ( c + v ) ,
showing that, for D > 2 v L c , in the RLSE T ⇒ , is D -independent and the same as that of the standard linear Sagnac effect in Eq. ( 1 ). The same conclusion holds for the round-trip T ⇐ of the co-propagating photon.
A.3.2 Showing that in general, for the RLSE the round trips intervals T ⇐ and T ⇒ depend on D . For two counter-propagating photons, the standard RLSE provides the same D -independent result Δ T of the standard linear Sagnac effect.
When during the round-trip interval T the device C * keeps on the same track (lower or upper contour section), there is no problem in deriving (as shown earlier) for Δ T = T ⇐ T ⇒ the D -independent results ( 1 )( 3 ), which are the same as those that can be found in literature. The calculations become quite complicated when the device C * passes, e.g., from the lower to the upper track and we wish to find results in general. Since in the variant S-RLSE, the round-trip T ⇒ has its maximum when the photon covers the maximum length 2 L γ (occurring when D 0 = v L ( γ c ) ), we consider here this special case and perform the calculations for deriving the round-trip T ⇐ of the co-moving photon and determine Δ T of ( 3 ). For the co-moving photon, we have first to find its position at t out = L ( γ c ) when point A reaches C * and the rod changes the direction of motion. With reference to Figure A3 (a), starting from C * and moving to the left at speed c toward point A, the photon reaches A when D 0 + v t = c t , i.e. ,
t A = D 0 c + v = v L γ c ( c + v ) .
After reaching A ( Figure A3 (b)), the photon moves to the upper track and, in the remaining interval, Δ t = t out t A propagates on the upper track toward point B while the rod is still moving to the right at speed v . Then, since A is moving at speed v , after the interval Δ t , the position of the photon relative to A is,
h = ( c v ) Δ t = ( c v ) ( t out t A ) = ( c v ) L γ ( c + v ) ,
as shown in Figure A3 (c).
Figure A3 Relative positions of photon and clock C * {{\rm{C}}}^{* } during light co-propagation in the round-trip interval T ⇐ {T}_{\Leftarrow } .
Figure A3
Relative positions of photon and clock C * during light co-propagation in the round-trip interval T ⇐ .
With the photon at distance h from A and the device C * coinciding with A and displaying the reading t out , the contour changes direction of motion. Then, from h , the photon reaches B when, h + c t = L γ v t , i.e. , after the interval,
t 2 = L γ h c + v = 1 c + v [ L γ ( c v ) L γ ( c + v ) ] = 2 v L γ ( c + v ) 2 = 2 v γ L ( 1 v c ) c ( c + v ) ,
as indicated in Figure A3 (d).
After reaching B on the upper track, the photon passes to the lower track and moves toward A, reaching it when L γ c t = v t at,
t 3 = L γ ( c v ) ,
as in shown Figure A3 (e).
At this point, the photon at A goes to the upper track and starts moving toward C * . However, in the interval t 2 + t 3 , point A has moved to the distance v ( t 2 + t 3 ) to the left of C * . Then, from A, the photon reaches C* when c t = v ( t 2 + t 3 ) after the interval,
t 4 = v ( t 2 + t 3 ) c = v L ( 2 v ( c v ) + ( c + v ) 2 ) γ c ( c + v ) 2 ( c v ) = v γ L ( c 2 + 4 v c v 2 ) c ( c + v ) c 2 .
After some algebra, the resulting round-trip time interval for the co-propagating photon is found to be,
(A16) T ⇐ = t out + t 2 + t 3 + t 4 = L γ c + 2 v γ L ( 1 v c ) c ( c + v ) + L γ ( c v ) + v γ L ( c 2 + 4 v c v 2 ) c ( c + v ) c 2 .
The sum t out + t 3 can be expressed as follows:
(A17) t out + t 3 = L γ c + L γ ( c v ) = 2 L γ c L γ c + L γ ( c v ) = 2 L γ c + v L γ c 2 ( 1 v c ) = 2 L γ c + γ v L ( 1 + v c ) c 2 .
By substituting Eq. ( A17 ) in Eq. ( A16 ), we find,
(A18) T ⇐ = 2 L γ c + γ v L ( 1 + v c ) c 2 + 2 v γ L ( 1 v c ) c ( c + v ) + v γ L ( c 2 + 4 v c v 2 ) c ( c + v ) c 2 = 2 L γ c + γ v L ( 4 c 2 + 4 c v ) c ( c + v ) c 2 = 2 L γ c + 4 γ v L c 2 ,
and with T ⇒ = 2 L ( γ c ) ,
Δ T = T ⇐ T ⇒ = 4 γ v L c 2
in agreement with the result of the standard linear Sagnac effect.
References
[1] Sagnac G. Regarding the proof for the existence of a luminiferous ether using a rotating interferometer experiment. C R Acad Sci. 1913;157:70810. Search in Google Scholar
[2] Wang R, Zhengb Y, Yao A, Langley D. Modified Sagnac experiment for measuring travel-time difference between counter-propagating light beams in a uniformly moving fiber. Phys Lett A. 2003;312:710. Wang R, Zheng Y, Yao A. Generalized Sagnac effect. Phys Rev Lett. 2004;93(14):143901. 10.1016/S0375-9601(03)00575-9 Search in Google Scholar
[3] Post EJ. Sagnac effect. Rev Mod Phys. 1967;39(2):47593. 10.1103/RevModPhys.39.475 Search in Google Scholar
[4] Lee C. Simultaneity in cylindrical spacetime. Am J Phys. 2020;88:131. 10.1119/10.0000002 Search in Google Scholar
[5] Klauber RD. Comments regarding recent articles on relativistically rotating frames. Am J Phys. 1999;67(2):1589. 10.1119/1.19213 Search in Google Scholar
[6] Selleri F. Noninvariant one-way velocity of light. Found Phys. 1996;26:641. Noninvariant one-way speed of light and locally equivalent reference frames. Found Phys Lett. 1997;10:7383. 10.1007/BF02764121 Search in Google Scholar
[7] Spavieri G, Gillies GT, GaarderHaug E, Sanchez A. Light propagation and local speed in the linear Sagnac effect. J Modern Optics. 2019;66(21):213141. 10.1080/09500340.2019.1695005 . Spavieri G, Gillies GT, Gaarder Haug E. The Sagnac effect and the role of simultaneity in relativity theory. J Mod Opt. 2021. doi: 10.1080/09500340.2021.1887384. Spavieri G. On measuring the one-way speed of light. Eur Phys J D. 2012;66:76. doi: 10.1140/epjd/e2012-20524-8; Spavieri G. Light propagation on a moving closed contour and the role of simultaneity in special relativity. Eur J Appl Phys. 2021;3:4:48. doi :10.24018/ejphysics.2021.3.4.99; Spavieri G, Gaarder Haug E. Testing light speed invariance by measuring the one-way light speed on earth. Physics Open 2022;12:100113. doi: 10.1016/j.physo.2022.100113. Search in Google Scholar
[8] Gift SJG. On the Selleri transformations: analysis of recent attempts by Kassner to resolve Selleris paradox. Appl Phys Res. 2015;7(2):112. 10.5539/apr.v7n2p112 Search in Google Scholar
[9] Kipreos ET, Balachandran RS. An approach to directly probe simultaneity. Modern Phys Lett A. 2016;31(26):1650157; Assessment of the relativistic rotational transformations. Modern Physics Letters A. 2021;36(16):2150113. Search in Google Scholar
[10] Landau LD, Lifshitz EML. The classical theory of fields. Vol. 2. 2nd English edn. Pergamon Press; 1962. p. 236. Search in Google Scholar
[11] Lundberg R. Critique of the Einstein clock variable. Phys essays. 2019;32:237; Travelling light. J Mod Opt. 2021;68(14). doi: 10.1080/09500340.2021.1945154. Search in Google Scholar
[12] Field JH. The Sagnac effect and transformations of relative velocities between inertial frames. Fund J Modern Phys. 2017;10(1):130. Search in Google Scholar
[13] Mansouri R, Sexl RU. A test theory of special relativity: I. Simultaneity and clock synchronization. Gen Rel Grav. 1977;8:497, 515, 809. 10.1007/BF00762634 Search in Google Scholar
[14] Schreiber KU, Gebauer A, Igel H, Wassermann J, Hurst RB, Wells JPR. The centennial of the Sagnac experiment in the optical regime: from a tabletop experiment to the variation of the Earths rotation. C R Physique. 2014;15:85965. http://dx.doi.org/10.1016/j.crhy.2014.10.003 . Search in Google Scholar
[15] Stedman GE. Ring-laser tests of fundamental physics and geophysics. Rep Prog Phys. 1997;60:615. 10.1088/0034-4885/60/6/001 Search in Google Scholar
[16] Malykin GB. The Sagnac effect: correct and incorrect explanations. Phys Uspekhi. 2000;43(12):122952. 10.1070/PU2000v043n12ABEH000830 Search in Google Scholar
[17] Ludlow DA, Boyd MM, Ye J, Peik E, Schmidt PO. Optical atomic clocks. Rev Mod Phys. 2015;87:637. 10.1103/RevModPhys.87.637 Search in Google Scholar
[18] Kim J, Chen J, Cox J, Kärtner FX. Attosecond-resolution timing jitter characterization of free-running mode-locked lasers. Optics Lett. 2007;32(24):351921; Kwon D, Jeon CG, Shin J, Heo MS, Park SE, Song Y, Kim J. Ultrafast, subnanometre-precision and multifunctional time-of-flight detection. Scientific Reports 2017:7:40917. Search in Google Scholar
[19] Tangherlini FR. An introduction to the general theory of relativity. Nuovo Cimento Suppl. 1961;20:1. 10.1007/BF02746778 Search in Google Scholar
[20] Spavieri G, Rodriguez M, Sanchez A. Thought experiment discriminating special relativity from preferred frame theories. J Phys Commun. 2018;2:085009. 10.1088/2399-6528/aad5fa . Search in Google Scholar
[21] de AbreuR, Guerra V. On the consistency between the assumption of a special system of reference and special relativity. Found Phys. 2006;36:182645. 10.1007/s10701-006-9085-5 Search in Google Scholar
[22] Bell JS. Speakable and unspeakable in quantum mechanics. Cambridge: Cambridge University Press; 1988. 10.1063/1.2811599 Search in Google Scholar
[23] Anderson R, Vetharaniam I, Stedman GE. Conventionality of synchronisation, gauge dependence and test theories of relativity. Phys Rep. 1998;295:93180. 10.1016/S0370-1573(97)00051-3 Search in Google Scholar
[24] Popper K. Conjectures and refutations. London: Routledge; 1963; Kuhn TS, The structure of scientific revolutions. Chicago, Illinois: University of Chicago Press; 1962. Search in Google Scholar
[25] Eisele C, Nevsky AY, Schiller S. Laboratory test of the isotropy of light propagation at the 1017 level. Phys Rev Lett. 2009;103:090401. 10.1103/PhysRevLett.103.090401 Search in Google Scholar PubMed
[26] Hughes VW, Robinson HG, Beltran-Lopez V. Upper limit for the anisotropy of inertial mass from nuclear resonance experiments. Phys Rev Lett. 1960;4:3424. 10.1103/PhysRevLett.4.342 Search in Google Scholar
[27] Drever RWP. A search for anisotropy of inertial mass using a free precession technique. Phil Mag. 1961;6:6837. 10.1080/14786436108244418 Search in Google Scholar
[28] Pruttivarasin T, Ramm M, Porsev SG, Tupitsyn II, Safronova MS, Hohensee MA, et al. Häffner H. Michelson-Morley analogue for electrons using trapped ions to test Lorentz symmetry. Nature. 2015;517:5925. 10.1038/nature14091 . Search in Google Scholar PubMed
[29] Thomas LH. The motion of the spinning electron. Nature (London). 1926;117:514; The kinematics of an electron with an axis. Phil Mag. 1927;3:122. Search in Google Scholar
[30] Jackson JD. Classical Electrodynamics, Sect. 11.8. 2nd edn. New York: John Wiley & Sons, Inc; 1975. Search in Google Scholar
[31] Bargmann V, Michel L, Telegdi VL. Precession of the polarization of particles moving in a homogeneous electromagnetic field. Phys Rev Lett. 1959;2:435. 10.1103/PhysRevLett.2.435 Search in Google Scholar
Received: 2023-06-08
Revised: 2023-08-21
Accepted: 2023-08-23
Published Online: 2023-09-16
© 2023 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
From the journal
Open Physics
Open Physics
Volume 21 Issue 1
Submit manuscript
Journal and Issue
This issue
All issues
Articles in the same Issue
Regular Articles
Dynamic properties of the attachment oscillator arising in the nanophysics
Parametric simulation of stagnation point flow of motile microorganism hybrid nanofluid across a circular cylinder with sinusoidal radius
Fractal-fractional advectiondiffusionreaction equations by Ritz approximation approach
Behaviour and onset of low-dimensional chaos with a periodically varying loss in single-mode homogeneously broadened laser
Ammonia gas-sensing behavior of uniform nanostructured PPy film prepared by simple-straightforward in situ chemical vapor oxidation
Analysis of the working mechanism and detection sensitivity of a flash detector
Flat and bent branes with inner structure in two-field mimetic gravity
Heat transfer analysis of the MHD stagnation-point flow of third-grade fluid over a porous sheet with thermal radiation effect: An algorithmic approach
Weighted survival functional entropy and its properties
Bioconvection effect in the Carreau nanofluid with CattaneoChristov heat flux using stagnation point flow in the entropy generation: Micromachines level study
Study on the impulse mechanism of optical films formed by laser plasma shock waves
Analysis of sweeping jet and film composite cooling using the decoupled model
Research on the influence of trapezoidal magnetization of bonded magnetic ring on cogging torque
Tripartite entanglement and entanglement transfer in a hybrid cavity magnomechanical system
Compounded Bell-G class of statistical models with applications to COVID-19 and actuarial data
Degradation of Vibrio cholerae from drinking water by the underwater capillary discharge
Multiple Lie symmetry solutions for effects of viscous on magnetohydrodynamic flow and heat transfer in non-Newtonian thin film
Thermal characterization of heat source (sink) on hybridized (CuAg/EG) nanofluid flow via solid stretchable sheet
Optimizing condition monitoring of ball bearings: An integrated approach using decision tree and extreme learning machine for effective decision-making
Study on the inter-porosity transfer rate and producing degree of matrix in fractured-porous gas reservoirs
Interstellar radiation as a Maxwell field: Improved numerical scheme and application to the spectral energy density
Numerical study of hybridized Williamson nanofluid flow with TC4 and Nichrome over an extending surface
Controlling the physical field using the shape function technique
Significance of heat and mass transport in peristaltic flow of Jeffrey material subject to chemical reaction and radiation phenomenon through a tapered channel
Complex dynamics of a sub-quadratic Lorenz-like system
Stability control in a helicoidal spinorbit-coupled open BoseBose mixture
Research on WPD and DBSCAN-L-ISOMAP for circuit fault feature extraction
Simulation for formation process of atomic orbitals by the finite difference time domain method based on the eight-element Dirac equation
A modified power-law model: Properties, estimation, and applications
Bayesian and non-Bayesian estimation of dynamic cumulative residual Tsallis entropy for moment exponential distribution under progressive censored type II
Computational analysis and biomechanical study of Oldroyd-B fluid with homogeneous and heterogeneous reactions through a vertical non-uniform channel
Predictability of machine learning framework in cross-section data
Chaotic characteristics and mixing performance of pseudoplastic fluids in a stirred tank
Isomorphic shut form valuation for quantum field theory and biological population models
Vibration sensitivity minimization of an ultra-stable optical reference cavity based on orthogonal experimental design
Effect of dysprosium on the radiation-shielding features of SiO 2 PbOB 2 O 3 glasses
Asymptotic formulations of anti-plane problems in pre-stressed compressible elastic laminates
A study on soliton, lump solutions to a generalized (3+1)-dimensional Hirota--Satsuma--Ito equation
Tangential electrostatic field at metal surfaces
Bioconvective gyrotactic microorganisms in third-grade nanofluid flow over a Riga surface with stratification: An approach to entropy minimization
Infrared spectroscopy for ageing assessment of insulating oils via dielectric loss factor and interfacial tension
Influence of cationic surfactants on the growth of gypsum crystals
Study on instability mechanism of KCl/PHPA drilling waste fluid
Analytical solutions of the extended KadomtsevPetviashvili equation in nonlinear media
A novel compact highly sensitive non-invasive microwave antenna sensor for blood glucose monitoring
Inspection of Couette and pressure-driven Poiseuille entropy-optimized dissipated flow in a suction/injection horizontal channel: Analytical solutions
Conserved vectors and solutions of the two-dimensional potential KP equation
The reciprocal linear effect, a new optical effect of the Sagnac type
Optimal interatomic potentials using modified method of least squares: Optimal form of interatomic potentials
The soliton solutions for stochastic CalogeroBogoyavlenskii Schiff equation in plasma physics/fluid mechanics
Research on absolute ranging technology of resampling phase comparison method based on FMCW
Analysis of Cu and Zn contents in aluminum alloys by femtosecond laser-ablation spark-induced breakdown spectroscopy
Nonsequential double ionization channels control of CO 2 molecules with counter-rotating two-color circularly polarized laser field by laser wavelength
Fractional-order modeling: Analysis of foam drainage and Fisher's equations
Thermo-solutal Marangoni convective Darcy-Forchheimer bio-hybrid nanofluid flow over a permeable disk with activation energy: Analysis of interfacial nanolayer thickness
Investigation on topology-optimized compressor piston by metal additive manufacturing technique: Analytical and numeric computational modeling using finite element analysis in ANSYS
Breast cancer segmentation using a hybrid AttendSeg architecture combined with a gravitational clustering optimization algorithm using mathematical modelling
On the localized and periodic solutions to the time-fractional Klein-Gordan equations: Optimal additive function method and new iterative method
3D thin-film nanofluid flow with heat transfer on an inclined disc by using HWCM
Numerical study of static pressure on the sonochemistry characteristics of the gas bubble under acoustic excitation
Optimal auxiliary function method for analyzing nonlinear system of coupled SchrödingerKdV equation with Caputo operator
Analysis of magnetized micropolar fluid subjected to generalized heat-mass transfer theories
Does the Mott problem extend to Geiger counters?
Stability analysis, phase plane analysis, and isolated soliton solution to the LGH equation in mathematical physics
Effects of Joule heating and reaction mechanisms on couple stress fluid flow with peristalsis in the presence of a porous material through an inclined channel
Bayesian and E-Bayesian estimation based on constant-stress partially accelerated life testing for inverted ToppLeone distribution
Dynamical and physical characteristics of soliton solutions to the (2+1)-dimensional KonopelchenkoDubrovsky system
Study of fractional variable order COVID-19 environmental transformation model
Sisko nanofluid flow through exponential stretching sheet with swimming of motile gyrotactic microorganisms: An application to nanoengineering
Influence of the regularization scheme in the QCD phase diagram in the PNJL model
Fixed-point theory and numerical analysis of an epidemic model with fractional calculus: Exploring dynamical behavior
Computational analysis of reconstructing current and sag of three-phase overhead line based on the TMR sensor array
Investigation of tripled sine-Gordon equation: Localized modes in multi-stacked long Josephson junctions
High-sensitivity on-chip temperature sensor based on cascaded microring resonators
Pathological study on uncertain numbers and proposed solutions for discrete fuzzy fractional order calculus
Bifurcation, chaotic behavior, and traveling wave solution of stochastic coupled KonnoOono equation with multiplicative noise in the Stratonovich sense
Thermal radiation and heat generation on three-dimensional Casson fluid motion via porous stretching surface with variable thermal conductivity
Numerical simulation and analysis of Airy's-type equation
A homotopy perturbation method with Elzaki transformation for solving the fractional BiswasMilovic model
Heat transfer performance of magnetohydrodynamic multiphase nanofluid flow of CuAl 2 O 3 /H 2 O over a stretching cylinder
ΛCDM and the principle of equivalence
Axisymmetric stagnation-point flow of non-Newtonian nanomaterial and heat transport over a lubricated surface: Hybrid homotopy analysis method simulations
HAM simulation for bioconvective magnetohydrodynamic flow of Walters-B fluid containing nanoparticles and microorganisms past a stretching sheet with velocity slip and convective conditions
Coupled heat and mass transfer mathematical study for lubricated non-Newtonian nanomaterial conveying oblique stagnation point flow: A comparison of viscous and viscoelastic nanofluid model
Power ToppLeone exponential negative family of distributions with numerical illustrations to engineering and biological data
Extracting solitary solutions of the nonlinear KaupKupershmidt (KK) equation by analytical method
A case study on the environmental and economic impact of photovoltaic systems in wastewater treatment plants
Application of IoT network for marine wildlife surveillance
Non-similar modeling and numerical simulations of microploar hybrid nanofluid adjacent to isothermal sphere
Joint optimization of two-dimensional warranty period and maintenance strategy considering availability and cost constraints
Numerical investigation of the flow characteristics involving dissipation and slip effects in a convectively nanofluid within a porous medium
Spectral uncertainty analysis of grassland and its camouflage materials based on land-based hyperspectral images
Application of low-altitude wind shear recognition algorithm and laser wind radar in aviation meteorological services
Investigation of different structures of screw extruders on the flow in direct ink writing SiC slurry based on LBM
Harmonic current suppression method of virtual DC motor based on fuzzy sliding mode
Micropolar flow and heat transfer within a permeable channel using the successive linearization method
Different lump k -soliton solutions to (2+1)-dimensional KdV system using Hirota binary Bell polynomials
Investigation of nanomaterials in flow of non-Newtonian liquid toward a stretchable surface
Weak beat frequency extraction method for photon Doppler signal with low signal-to-noise ratio
Electrokinetic energy conversion of nanofluids in porous microtubes with Greens function
Examining the role of activation energy and convective boundary conditions in nanofluid behavior of Couette-Poiseuille flow
Review Article
Effects of stretching on phase transformation of PVDF and its copolymers: A review
Special Issue on Transport phenomena and thermal analysis in micro/nano-scale structure surfaces - Part IV
Prediction and monitoring model for farmland environmental system using soil sensor and neural network algorithm
Special Issue on Advanced Topics on the Modelling and Assessment of Complicated Physical Phenomena - Part III
Some standard and nonstandard finite difference schemes for a reactiondiffusionchemotaxis model
Special Issue on Advanced Energy Materials - Part II
Rapid productivity prediction method for frac hits affected wells based on gas reservoir numerical simulation and probability method
Special Issue on Novel Numerical and Analytical Techniques for Fractional Nonlinear Schrodinger Type - Part III
Adomian decomposition method for solution of fourteenth order boundary value problems
New soliton solutions of modified (3+1)-D WazwazBenjaminBonaMahony and (2+1)-D cubic KleinGordon equations using first integral method
On traveling wave solutions to Manakov model with variable coefficients
Rational approximation for solving Fredholm integro-differential equations by new algorithm
Special Issue on Predicting pattern alterations in nature - Part I
Modeling the monkeypox infection using the MittagLeffler kernel
Spectral analysis of variable-order multi-terms fractional differential equations
Special Issue on Nanomaterial utilization and structural optimization - Part I
Heat treatment and tensile test of 3D-printed parts manufactured at different build orientations
Subjects
Architecture and Design
Arts
Asian and Pacific Studies
Business and Economics
Chemistry
Classical and Ancient Near Eastern Studies
Computer Sciences
Cultural Studies
Engineering
General Interest
Geosciences
History
Industrial Chemistry
Islamic and Middle Eastern Studies
Jewish Studies
Law
Library and Information Science, Book Studies
Life Sciences
Linguistics and Semiotics
Literary Studies
Materials Sciences
Mathematics
Medicine
Music
Pharmacy
Philosophy
Physics
Social Sciences
Sports and Recreation
Theology and Religion
Services
For Journal Authors
For Book Authors
For Librarians
Rights & Permissions
Publications
Publication types
Open Access
About
Contact
Career
About De Gruyter
Partnerships
Press
FAQs
Social
Facebook
Instagram
LinkedIn
X / Twitter
YouTube
De Gruyter
Open-Athens Winner of the OpenAthens
Best Publisher UX Award 2022
Help/FAQ
Privacy policy
Cookie Policy
Accessibility
Terms & Conditions
Legal Notice
© Walter de Gruyter GmbH 2024
Consent to website analysis
We use cookies and other technologies. Some of them are necessary for the website to function and are always set. Cookies for website analysis are not required and are set only with your consent. Some services for analysis process personal data in the USA. With your consent to use these services, you also consent to the processing of your data in the USA. Your consent is voluntary and can be revoked at any time. For more information, please see our Cookie Policy.
Accept optional analytics cookies Reject non-essential cookies