Abstract
The charged particle confined by a harmonic potential in a noncommutative planar phase space interacting with a homogeneous dynamical magnetic field and Aharonov-Bohm potentials is studied. We find that the canonical orbital angular momenta of the reduced models, which are obtained by setting the mass and a dimensionless parameter to zero, take fractional values. These fractional angular momenta are not only determined by the flux inside the thin long solenoid but also affected by the noncommutativities of phase space.
The Aharonov–Bohm (AB) effect exhibits the shift of the interference pattern in the double-slit experiment with a thin long solenoid located between these two slits [1]. It indicates that quantum mechanically, charged particles are influenced by magnetic vectors in the region where the field strengths are zero. It is quite different from the classical mechanics since the classical motion of the charged particle is not affected by the AB potentials. Therefore, AB effect reveals the fundamental roles the magnetic potentials play in quantum theories. It is well known that contrary to three-dimensional space in which angular momenta can only take half-integer values (we set ћ=1), angular momenta in a plane can take fractional values. It leads to an important concept in theoretical physics, namely, aynons [2], [3], [4]. There are many methods to realise the fractional angular momentum. One of the most classical methods of realising fractional angular momentum is to couple a charged planar particle to the Abelian Chern-Simons gauge field [2], [3], [4].
Recently, the author of [5] proposed an interesting method to realise the fractional angular momentum. He considered a planar charged particle (take, for example, an ion) confined by a harmonic potential in the background of a dynamical magnetic field[1] and AB magnetic vector potentials. The author found that the angular momentum of the reduced model, which is obtained by setting the kinetic energy to one of its eigenvalues, i.e. the lowest energy level, takes the fractional values. This fractional angular momentum is induced by the AB vector potentials and is proportional to the magnetic flux inside the thin long solenoid. The dynamical magnetic field plays an interesting role: although it does not contribute to the fractional angular momentum, the fractional zero-point angular momentum will not appear in the absence of it.
On the other hand, spatial non-commutativity attracts much attention nowadays because of the string theory [6], [7], [8]. In fact, it had a long history in physics [9], [10]. It is known that the spatial noncommutativity arises naturally from string theory [11], [12], [13], [14]. There are numerous papers about quantum field theories on the non-commutative space, including both perturbative and non-perturbative aspects [15], [16], [17].
Quantum mechanics on the noncommutative space has also been studied for both non-relativistic and relativistic models. The most popular method of studying quantum mechanics in noncommutative space is to map the noncommutative space to a commutative one by the Bopp shift (or the generalised Bopp shift) and then study them in the commutative space [18], [19], [20], [21], [22], [23]. Most exactly solvable models were studied by using this method. Path integral formulation in noncommutative mechanics has also been investigated [24]. Based on the path integral formulation in noncommutative space, the authors of Refs. [25], [26] deduced an effective Lagrangian and gave the noncommutative corrections to the AB effect in noncommutative space. The application of path integral formulation to solve the spectra of noncommutative quantum mechanical models has also been investigated [27], [28], [29].
In this paper, we shall generalise the work of Ref. [5] to the noncommutative planar phase space; i.e. both the coordinates and momenta are noncommutative simultaneously. We shall show that the fractional angular momenta can also be obtained in the noncommutative planar phase space if certain limits are taken.[2] The noncommutative plane is described by the commutators (the lowercase Latin indexes run from 1 to 2)
where θ is the noncommutative parameter between coordinates which is taken as a constant throughout this paper, and ∈ij is the two-dimensional anti-symmetrical tensor.
The noncommutative planar phase space is defined by the commutation relations
where η is the noncommutative parameter between momenta (it is also taken as a constant in this paper). Compared with the noncommutative plane (1), we can see that the momenta are also noncommutative in the noncommutative phase space. Quantum mechanical models on the noncommutative phase spaces which are characterised by the algebraic relations similar with (2) were widely studied in the past years [30], [31], [32], [33], [34], [35], [36], [37], [38], [39], [40], [41], [42], [43], [44], [45], [46], [47].
We can construct the Lagrangian which gives the classical version (i.e.
in which κ=1−θη, and H is a specific Hamiltonian. In this paper we shall focus on the harmonic potential. Thus, the Hamiltonian is
with K being a constant. Note that θ and η have the dimensions of L2 and L−2 , respectively, so κ is dimensionless. We should note that the noncommutativities, if they do exist, will be extremely small [31], [32], i.e. |θ|, |η|=1. Therefore, there is no singularity in the Lagrangian (3).
There are two different kinds of magnetic vector potentials we shall introduce. One kind is the dynamical magnetic vector potentials. The other is the AB ones which are realised by putting a thin long solenoid vertical to the plane. We introduce these magnetic vector potentials into the model (3) by the minimal substitution [48], [49], [50]
where Ai and
for the dynamical magnetic field with B being the field strength and
for the AB potentials outside the solenoid with Φ0 being the flux inside the thin long solenoid.
Thus, the model we are interested in the present study is described by the Lagrangian (outside the solenoid)
Substituting the expressions of magnetic vector potentials (6, 7) into the above Lagrangian and writing the term
where H has been given in (4) and
is the combination of the dynamical magnetic field and the noncommutative parameter between momenta. Clearly, the noncommutative parameter between momenta contributes to the dynamical magnetic field.
We will show that the model (9) has two different reduced models which are obtained by setting the mass and a dimensionless parameter to zero. The angular momentum of each reduced model takes fractional values. Therefore, there are two different mechanisms to get the fractional angular momentum from the model (9).
We shall quantise the model (9) canonically before studying its quantum properties. In doing so, we introduce the canonical momenta with respects to the variables (xi, pi). They are
The basic non-vanishing Poisson brackets among canonical variables
The canonical orbital angular momentum, by definition, is
The introduction of canonical momenta (11) does not contain ‘velocity’ terms. It results in algebraic relations among canonical variables
The primary constraints are
in which ‘≈’ means equivalent on the constraint hypersurface. For future convenience, we label the primary constraints
The Poisson brackets among the primary constraints ΦI are
The determinant of this matrix is
in which
is also a dimensionless parameter.
Clearly, the determinant of the matrix {ΦI, ΦJ} can be modulated by adjusting the magnitude of the dynamical magnetic field. It will vanish when the dimensionless parameter γ takes zero value (or when the magnitude of the dynamical magnetic field takes critical value
Thus, we shall investigate the cases of Det{ΦI, ΦJ}≠0 and Det{ΦI, ΦJ}=0 (or, equivalently, cases of γ≠0 and γ=0), respectively, in the following. We shall show that there are different mechanisms in both cases to get the fractional angular momenta. Let us study the case of γ≠0 firstly. In this case, the determinant of the matrix {ΦI, ΦJ} does not vanish. The inverse matrix of {ΦI, ΦJ} is
The consistency conditions of primary constraints are given by
in which λI are Lagrange multipliers. Substituting the Hamiltonian (4) and the explicit expressions of constraints ΦI into the above equation, we find that the consistency conditions only determine the Lagrangian multipliers λI. Therefore, in the case of γ≠0, ΦI≈0 (I=1, 2, 3, 4) exhausts all the constraints of the model (9), and they belong to the second class [51].
We must get the Dirac brackets among variables
where F, G are two arbitrary functions of the canonical variables
The quantum commutators are obtained by the replacement [ , ]→i{ , }D [51].
The constraints ΦI≈0 (I=1, 2, 3, 4) are second class; they are ‘strong’ equations and can be used to eliminate the redundant degrees of freedom in the angular momentum (13). Thus, we write the angular momentum (13) in terms of (xi, pi)
It can be checked directly by using the commutators [the quantum version of the Dirac brackets (22)] that
and
It shows that the angular momentum (23) is conserved and is the generator of rotation.
In Ref. [5], the author proposed an interesting method to obtain the fractional angular momentum from the commutative counterpart (θ=η=0) of (9). He found that the angular momentum of the reduced model, which was attained by setting the eigenvalues of the kinetic energy to the lowest energy level, took fractional values. This fractional value is proportional to the flux inside the thin long solenoid. The dynamical magnetic field plays an interesting role: although the dynamical magnetic field does not contribute to the fractional angular momentum, the fractional angular momentum does not appear in the absence of it.
We shall show that the fractional angular momenta can also be obtained from the noncommutative model (9) if certain limits are chosen. In fact, there are two different mechanisms to get the fractional angular momenta in the model (9).
Let us consider the first mechanism of getting the fractional angular momentum. It is to set the mass to zero, i.e. m→0 in the model (9). The zero-mass limit was first considered in Ref. [52] during the studies of Chern-Simons quantum mechanics. Then, it was generalised to noncommutative cases in Refs. [27], [29], [53].
In view of the Hamiltonian (4), we have to set
so as to avoid the divergency of the Hamiltonian. It means that the kinetic energy is neglected.
The Lagrangian (9) reduces to the form
when the limit m→0 is taken.
Therefore, besides the primary constraints ΦI≈0 (I=1, 2, 3, 4), there are two extra constraints, namely,
We label all the constraints as Ψα=(ΦI, χi)≈0 (α=1, 2, …, 6). It can be verified that the constraints Ψα≈0 are second class and there are no further constraints.
Since the constraints χi=pi≈0 are second class, they can be used to eliminate the redundant degrees of freedom. As a result, the angular momentum (23) is further simplified to the form
We must know the commutation relations between variables xi before calculating the eigenvalues of the angular momentum (29). To this end, we shall calculate the Dirac brackets between xi.
The explicit expression of the matrix {Ψα, Ψβ} is
The inverse of this matrix is given by
After some direct calculations, we get
Therefore, the commutation relations among the variables xi are
In order to express the angular momentum (29) lucidly, we introduce a new pair of variables (X, P) which satisfy [X, P]=i. They are
In terms of variables (X, P), we rewrite the angular momentum (29) in the form
By considering the commutation relation between X and P, we find that the first part in angular (35) is nothing but the Hamiltonian of a one-dimensional harmonic oscillator with unit mass and frequency. Thus, its eigenvalues can be written down directly. They are
Obviously, there is an extra term in the eigenvalues of the angular momentum which is proportional to the flux inside the solenoid. It can take fractional values.
We have to emphasise that there is an exception in the case of γ≠0. To illustrate it clearly, let us consider the situation BT=0. It implies that the dynamical magnetic field cancels the effect of noncommutativity between the momenta.
The Lagrangian (27) is further simplified to the form
when BT=0.
We introduce the canonical momenta with respective to xi
Again, the introduction of the canonical momenta leads to primary constraints. We label them as
Different from the situation BT≠0 in which there are no secondary constraints, the consistency conditions of the primary constraints
The constraints pi≈0 and xi≈0 means that there are no dynamics. Therefore, we cannot get the fractional angular momentum.
In the following, we shall analyse the other mechanism to get the fractional angular momentum from the model (9). We shall show that the fractional angular momenta will arise naturally when the dimensionless parameter γ=0 (or, equivalently, when the magnetic field takes the value
In the case of γ=0, the matrix (16) is singular. We should verify whether there are secondary constraints. The consistency conditions of the primary constraints are
where μi, νi are Lagrange multipliers. These consistency conditions lead to
The above equations determine the Lagrange multipliers μi and νi uniquely if γ≠0. Therefore, there are no secondary constraints in the case of γ≠0. However, when the parameter γ takes zero value, the above equations will lead to secondary constraints. It is straightforward to check that (43) and (44) are not independent when γ=0. They are equivalent up to an overall constant. Thus, we choose
as the secondary constraints.
The consistency conditions of the secondary constraints (45) do not lead to further constraints. And the constraints (45) are also second class. They are the ‘strong’ equation and can be used to eliminate the redundant variables in the expression of angular momentum. Thus, the angular momentum (23) becomes
We must determine the commutation relation among the variables xi in order to get the eigenvalues of the angular momentum (46). The calculation of the Dirac brackets among xi is unavoidable.
There are six constraints in total. They are
The inverse matrix is given by
The Dirac brackets among xi can be derived by using the matrix (48) directly. They are
So the commutation relations among the variables xi become
We introduce a pair of variables
The angular momentum (46) becomes
in terms of variables
It shows that when the dimensionless parameter γ takes zero value (or when the magnetic field takes the value
Compared with the result of the commutative counterpart (θ, η→0) in Ref. [5], we find that the fractional angular momenta in the noncommutative model (9) are modified by the noncommutative parameter η. As a consistency check, it is easy to see that our result (36) is identical to the one in Ref. [5] when the noncommutative parameters θ, η take zero values.
To summarise, we generalise the studies in Ref. [5] to the planar noncommutative phase space. we find that there are two different reduced models of (9). They are obtained by setting the mass and the dimensionless parameter γ=1−θBT to zero from (9), respectively. We pay our attention to the angular momenta of the reduced models. We find that the angular momentum of each reduced model takes fractional value. Compared with its commutative counterpart [5], we find that the fractional angular momentum is not only proportional to the flux inside the thin long solenoid but also affected by the noncommutativities of coordinates and momenta.
Acknowledgements
This work is supported by the NSFC with Grant No 11465006.
References
[1] Y. Aharonove and D. Bohm, Phys. Rev. 115, 485 (1959).10.1103/PhysRev.115.485Suche in Google Scholar
[2] F. Wilczek, Fractional Statistics and Anyon Superconditivity, World Scientific, Singapore 1990.10.1142/0961Suche in Google Scholar
[3] A. Lerda, Anyons, Quantum Mechanics of Particles with Fractional Statistics, Lectures Notes in Physics, Springer-Verlag, Berlin 1992.10.1007/978-3-540-47466-1Suche in Google Scholar
[4] A. Khare, Fractional Statistics and Quantum Theory, World Scientific, Singapore 1997.10.1142/2988Suche in Google Scholar
[5] J. Z. Zhang, Phys. Lett. B 670, 205 (2008).10.1016/j.physletb.2008.10.056Suche in Google Scholar
[6] A. Connes, M. Douglas, and A. S. Schwarz, J. High Energy Phys. 02, 003 (1998).10.1088/1126-6708/1998/02/003Suche in Google Scholar
[7] N. Seiberg and E. Witten, J. High Energy Phys. 09, 032 (1999).10.1088/1126-6708/1999/09/032Suche in Google Scholar
[8] M. R. Douglas and N. A. Nekrasov, Rev. Mod. Phys. 73, 977 (2001).10.1103/RevModPhys.73.977Suche in Google Scholar
[9] H. S. Snyder, Phys. Rev. 71, 38 (1946).10.1103/PhysRev.71.38Suche in Google Scholar
[10] C. N. Yang, Phys. Rev. 71, 874 (1947).10.1103/PhysRev.72.874Suche in Google Scholar
[11] C. S. Chu and P. M. Ho, Nucl. Phys. B 550, 151 (1999).10.1016/S0550-3213(99)00199-6Suche in Google Scholar
[12] C. S. Chu and P. M. Ho, Nucl. Phys. B 568, 447 (2000).10.1016/S0550-3213(99)00685-9Suche in Google Scholar
[13] F. Ardalan, H. Arfaei, and M. M. Sheikh-Jabbari, Nucl. Phys. B 576, 578 (2000).10.1016/S0550-3213(00)00096-1Suche in Google Scholar
[14] J. Jing and Z.-W. Long, Phys. Rev. D 72, 126002 (2005).10.1103/PhysRevD.72.126002Suche in Google Scholar
[15] S. Minwalla, M. Van Raamsdonk, and N. Seiberg, J. High Energy Phys. 02, 020 (2000).10.1088/1126-6708/2000/02/020Suche in Google Scholar
[16] M. Van Raamsdonk and N. Seiberg, J. High Energy Phys. 03, 035 (2000).10.1088/1126-6708/2000/03/035Suche in Google Scholar
[17] R. Gopakumar, S. Minwalla, and A. Strominger, J. High Energy Phys. 05, 020 (2000).10.1088/1126-6708/2000/05/020Suche in Google Scholar
[18] M. Chaichian, M. M. Sheikh-Jabbari, and A. Tureanu, Phys. Rev. Lett. 86, 2716 (2001).10.1103/PhysRevLett.86.2716Suche in Google Scholar
[19] B. Morariu and A. P. Polychronakos, Nucl. Phys. B 610, 531 (2001).10.1016/S0550-3213(01)00294-2Suche in Google Scholar
[20] V. P. Nair and A. P. Polychronakos, Phys. Lett. B 505, 267 (2001).10.1016/S0370-2693(01)00339-2Suche in Google Scholar
[21] S. Bellucci, A. Nersessian, and C. Sochichiu, Phys. Lett. B 522, 345 (2001).10.1016/S0370-2693(01)01304-1Suche in Google Scholar
[22] D. Karabali, V. P. Nair, and A. P. Polychronakos, Nucl. Phys. B 627, 565 (2002).10.1016/S0550-3213(02)00062-7Suche in Google Scholar
[23] B. Morariu and A. P. Polychronakos, Nucl. Phys. B 634, 326 (2002).10.1016/S0550-3213(02)00298-5Suche in Google Scholar
[24] C. Acatrinei, J. High Energy Phys. 09, 007 (2001).10.1088/1126-6708/2001/09/007Suche in Google Scholar
[25] M. Chaichian, P. Presnajder, M. M Sheikh-Jabbari, and A. Tureanu, Phys. Lett. B 527, 149 (2002).10.1016/S0370-2693(02)01176-0Suche in Google Scholar
[26] M. Chaichian, A. Demichev, P. Presnajder, M. M Sheikh-Jabbari, and A. Tureanu, Nucl. Phys. B 611, 383 (2001).10.1016/S0550-3213(01)00348-0Suche in Google Scholar
[27] J. Jing, F. H. Liu, and J. F. Chen, Phys. Rev. D 78, 125004 (2008).10.1103/PhysRevD.78.125004Suche in Google Scholar
[28] J. Jing, J. Y. Ma, Y. H. Guo, and J. F. Chen, Euro. Phys. J. C 56, 591 (2008).10.1140/epjc/s10052-008-0682-7Suche in Google Scholar
[29] J. Jing, F. H. Liu, and J. F. Chen, Europhys. Lett. 84, 61001 (2008).10.1209/0295-5075/84/61001Suche in Google Scholar
[30] C. Bastos, O. Bertolami, N. C. Dias, and J. N. Prata, Phys. Rev. D 86, 105030 (2012).10.1103/PhysRevD.86.105030Suche in Google Scholar
[31] O. Bertolami, J. G. Rosa, C. M. L. de Aragao, P. Castorina, and D. Zappala, Phys. Rev. D 72, 025010 (2005).10.1103/PhysRevD.72.025010Suche in Google Scholar
[32] C. Bastos and O. Bertolami Phys. Lett. A 372, 5556 (2008).10.1016/j.physleta.2008.06.073Suche in Google Scholar
[33] C. Bastos, O. Bertolami, N. C. Dias, and J. N. Prata, J. Math. Phys. 49, 072101 (2008).10.1063/1.2944996Suche in Google Scholar
[34] C. Bastos, O. Bertolami, N. C. Dias, and J. N. Prata, Phys. Rev. D 78, 023516 (2008).10.1103/PhysRevD.78.023516Suche in Google Scholar
[35] C. Bastos, O. Bertolami, N. C. Dias, and J. N. Prata, Phys. Rev. D 80, 124038 (2009).10.1103/PhysRevD.80.124038Suche in Google Scholar
[36] C. Bastos, O. Bertolami, N. C. Dias, and J. N. Prata, Phys. Rev. D 82, 041502 (2010).10.1103/PhysRevD.82.041502Suche in Google Scholar
[37] C. Bastos, O. Bertolami, N. C. Dias, and J. N. Prata, Phys. Rev. D 84, 024005 (2011).10.1103/PhysRevD.84.024005Suche in Google Scholar
[38] O. Bertolami and Carlos. A. D. Zarro, Phys. Rev. D 81, 025005 (2010).10.1103/PhysRevD.81.025005Suche in Google Scholar
[39] A. E. Bernardini and O. Bertolami, Phys. Rev. A 88, 012101 (2013).10.1103/PhysRevA.88.012101Suche in Google Scholar
[40] C. Bastos, A. E. Bernardini, O. Bertolami, N. Dias, and J. Prata, Phys. Rev. D 90, 045023 (2014).10.1103/PhysRevD.90.045023Suche in Google Scholar
[41] C. Bastos, A. E. Bernardini, O. Bertolami, N. Dias, and J. Prata, Phys. Rev. D 91, 065036 (2015).10.1103/PhysRevD.91.065036Suche in Google Scholar
[42] J. Z. Zhang, Ann. Phys. 324, 1847 (2009).10.1016/j.aop.2009.06.002Suche in Google Scholar
[43] J. Z. Zhang, Phys. Rev. D 74, 124005 (2006).10.1103/PhysRevD.74.124005Suche in Google Scholar
[44] J. Z. Zhang, K. L. Gao, and C. G. Ning, Phys. Rev. D 78, 105201 (2008).10.1103/PhysRevD.78.105021Suche in Google Scholar
[45] J. Z. Zhang, Phys. Lett. B 597, 362 (2004).10.1016/j.physletb.2004.07.039Suche in Google Scholar
[46] Q. J. Yin and J. Z. Zhang, Phys. Lett. B 613, 91 (2005).10.1016/j.physletb.2005.03.040Suche in Google Scholar
[47] J. Z. Zhang, Phys. Lett. B 639, 403 (2006).10.1016/j.physletb.2006.06.040Suche in Google Scholar
[48] C. Duval and P. A. Horvathy, Phys. Lett. B 479, 284 (2000).10.1016/S0370-2693(00)00341-5Suche in Google Scholar
[49] C. Duval and P. A. Horvathy, J. Phys. A 34, 10097 (2001).10.1088/0305-4470/34/47/314Suche in Google Scholar
[50] P. A. Horvathy, Ann. Phys. (N.Y.) 299, 128 (2002).10.1006/aphy.2002.6271Suche in Google Scholar
[51] P. A. M. Dirac, Lecture Notes on Quantum Mechanics, Yeshiva Uinversity, NewYork 1964.Suche in Google Scholar
[52] G. V. Dunne, R. Jackiw, and C. A. Trugenberger, Phys. Rev. D 41, 661 (1990).10.1103/PhysRevD.41.661Suche in Google Scholar PubMed
[53] T. Jiang, Z. G. Yuan, Q. Wang, Z. W. Long, and J. Jing, Int. J. Mod. Phys. A 27, 1250092 (2012).10.1142/S0217751X12500923Suche in Google Scholar
©2016 Walter de Gruyter GmbH, Berlin/Boston
Artikel in diesem Heft
- Frontmatter
- A Few New 2+1-Dimensional Nonlinear Dynamics and the Representation of Riemann Curvature Tensors
- First-Principle Study of the Structural, Electronic, and Optical Properties of Cubic InNxP1–x Ternary Alloys under Hydrostatic Pressure
- Bäcklund Transformation and Soliton Solutions for a (3+1)-Dimensional Variable-Coefficient Breaking Soliton Equation
- Analytical Solitons for Langmuir Waves in Plasma Physics with Cubic Nonlinearity and Perturbations
- Generalized Klein-Gordon and Dirac Equations from Nonlocal Kinetic Approach
- Fractional Zero-Point Angular Momenta in Noncommutative Quantum Mechanics
- A Crossover from High Stiffness to High Hardness: The Case of Osmium and Its Borides
- Impact of Entropy Generation on Stagnation-Point Flow of Sutterby Nanofluid: A Numerical Analysis
- High-Frequency Waves in a Random Distribution of Metallic Nanoparticles in an External Magnetic Field
- Differential Invariants of the (2+1)-Dimensional Breaking Soliton Equation
- Asymptotic Analysis to Two Nonlinear Equations in Fluid Mechanics by Homotopy Renormalisation Method
- Massive Particle Reflection from Moving Mirrors
Artikel in diesem Heft
- Frontmatter
- A Few New 2+1-Dimensional Nonlinear Dynamics and the Representation of Riemann Curvature Tensors
- First-Principle Study of the Structural, Electronic, and Optical Properties of Cubic InNxP1–x Ternary Alloys under Hydrostatic Pressure
- Bäcklund Transformation and Soliton Solutions for a (3+1)-Dimensional Variable-Coefficient Breaking Soliton Equation
- Analytical Solitons for Langmuir Waves in Plasma Physics with Cubic Nonlinearity and Perturbations
- Generalized Klein-Gordon and Dirac Equations from Nonlocal Kinetic Approach
- Fractional Zero-Point Angular Momenta in Noncommutative Quantum Mechanics
- A Crossover from High Stiffness to High Hardness: The Case of Osmium and Its Borides
- Impact of Entropy Generation on Stagnation-Point Flow of Sutterby Nanofluid: A Numerical Analysis
- High-Frequency Waves in a Random Distribution of Metallic Nanoparticles in an External Magnetic Field
- Differential Invariants of the (2+1)-Dimensional Breaking Soliton Equation
- Asymptotic Analysis to Two Nonlinear Equations in Fluid Mechanics by Homotopy Renormalisation Method
- Massive Particle Reflection from Moving Mirrors