Quantum Commnucation Note
Quantum Commnucation Note
Sougato Bose
January 11, 2021
1
Figure 1: Schematic examples of bits used in classical computers. Note their
robustness – if a single or a few spins flip then nothing much happens to the
net magnetic field generated and hence to the value of the bit; Similar holds
for small amounts of charge shifting from one plate of the capacitor to the
other.
Figure 2: The ultimate miniaturized form of bits: a single spin (of an atom
or an electron or a nuclear spin) or a single charge in one box or other –
a double quantum dot. These single systems are also quantum mechanical
systems
2
representation of the state of bits as column vectors with two components is
that this representation is more versatile than a single number and will also
enable us to represent all the states of a qubit. Once bits are represented as
column vectors as in the above Eq.(1), logic gates have to be represented as
matrices. For example, the NOT gate1 that changes 0 to 1 and vice versa is
given by the matrix
0 1
NOT = . (2)
1 0
We will often call it the X gate (as is the norm in the field of quantum
information, where this gate is also used). As far as operations on a single
bit are concerned, you can do only two things: (a) not do anything, i.e. the
1 0
Identity gate I = and (b) flip the bit, i.e., the NOT (or X) gate.
0 1
However, when more bits are available, there are further operations that one
may do. For ascertaining the matrix representation of these gates, one first
has to specify the column vector representation of states of multiple bits.
For two bits, for example, one can represent the states as one of four options
00, 01, 10 and 11, and thereby by a column matrix of four components as
follows
2 3 2 3 2 3 2 3
1 0 0 0
6 0 7 6 1 7 6 0 7 6 0 7
00 ⌘ 6 7 6 7 6 7 6 7
4 0 5 , 01 ⌘ 4 0 5 , 10 ⌘ 4 1 5 , 11 ⌘ 4 0 5 . (3)
0 0 0 1
Note that the above way of writing can be generalized to many bits simply
by noting that the
vector form of 00 can be generated by taking the outer
1 1
product ⌦ . For example, following this convention, the vector
0 0
representation of the bit string 001011 is
1 1 0 1 0 0
001011 = ⌦ ⌦ ⌦ ⌦ ⌦ (4)
0 0 1 0 1 1
Exercise .1
Write down the 8 component column vector representing to the state 010 of
3 bits.
We can now represent some logic gates on two bits using matrices. For
example the XOR gate (which we will call a CNOT gate, as this is its name in
the field of quantum information) accomplishes the following transformation
1
We assume familiarity with basic boolean logic gates.
3
on two bits: 00 ! 00, 01 ! 01, 10 ! 11, 11 ! 10, and can thereby be
represented by the matrix
2 3
1 0 0 0
6 0 1 0 0 7
CNOT = 6
4 0 0 0 1 5.
7 (5)
0 0 1 0
The trick for writing matrices in the above form is that you arrange
00, 01, 10 and 11 along rows and columns along the top and the left of the
matrix, and then write 1 in the places to which a certain element goes (e.g.,
as 10 goes to 11, the (4,3) element is 1, while the rest of the elements in
column 3 are 0..
Exercise .2
Write down the 8 ⇥ 8 matrix for the Controlled-Controlled-NOT (or TOF-
FOLI) Gate.
Exercise .3
Write down the 8 ⇥ 8 matrix for a Gate P that permutes bits cyclically, i.e.,
P |abci = |cabi.
Exercise .4
Show that the gates CNOT2 = I, TOFFOLI2 = I, but P 2 6= I, and instead
P P = I and P 3 = I.
T
4
a
ab
ab
ab
ab
b
Figure 3: The figure shows how a simple physical process can be regarded as
a computation.
5 Qubits
We now proceed to describe the peculiar behaviour that the tiniest possible
magnetic moments in nature, the so called spin-1/2 particles. These would be
the ultimate miniaturized version of the magnetic domain bits, where each
domain is now replaced by a single particle carrying the smallest possible
magnetic moment (an electron is an example of such a spin-1/2 particle,
as are several types of atoms). As will become clear to you soon, their
behaviour in experiments readily show that they violate the property (a) of
bits mentioned in the begining of the chapter – therefore it is not right to
call them bits any more – rather they are a classic example of a qubit.
One may measure the magnitude of the magnetic moment of the spin-1/2
particle (henceforth only referred to as a spin) by the famous Stern-Gerlach
5
experiment where the particle is made to fly through an inhomogeneous mag-
netic field oriented along the direction in which one wants to measure the
magnetic moment. For example, Fig.4 shows a Stern Gerlach apparatus for
measuring the magnetic moment of a particle along the Z (vertical) direction.
The magnetic moment is measured from the direction towards which its path
bends while leaving the inhomogeneous magnetic field – for example, if an
upwards deflection from the original direction signifies a positive magnetic
moment then a downwards deflection signifies a negative magnetic moment.
The magnitude of the deflection (i.e., the angle by which it deflects) gives the
magnitude of the magnetic moment. The most striking result here is that
measuring the value of the the magnetic moment of a spin-1/2 particle in any
direction yields only two values which can be labelled as +1 (corresponding
to a magnetic moment +µB in relevant units) and -1 (corresponding to µB
in relevant units). No other value of magnetic moment, apart from ±µB , is
ever observed in any experiment, though the direction of the initial magnetic
moment can be anything! This experimental fact is very peculiar and readily
implies two important things:
1. The magnetic moment directions of the spin (and indeed the state of
any quantum mechanical system) can randomly change (or what we shall
call jump) when measured, manifestly violating the fact that a bit can be
measured without changing its state.
Let us now make the case that there are sudden jumps induced by a mea-
surement a bit clearer and also give you the probabilties of these jumps as has
been deduced by experimentation. For example, the Stern-Gerlach apparatus
along the Z direction of Fig.4 could have been preceded by another Stern-
Gerlach apparatus in any other direction such as a direction making an angle
✓ with the Z direction. When a spin exiting that preceding apparatus with a
magnetic moment +1 is observed, then we are sure that the spin entering our
Stern-Gerlach apparatus oriented along the Z direction possess a magnetic
moment along a direction ✓ to the Z axis. Yet when it exits the Z oriented
Stern-Gerlach apparatus, its magnetic moment is found to be (measured to
be) exclusively either along +Z or Z directions. Thereby the magnetic
moment must be making a jump from the direction ✓ with the Z axis to a di-
rection +Z or Z by virtue of passing through the Z oriented Stern-Gerlach
apparatus. Moreover the probablility of these jumps as dictated by exper-
imental observations are found to be cos2 ✓/2 and sin2 ✓/2 = cos2 (✓ + ⇡)/2
to jump to the +Z and Z axes respectively. Therefore the general rule for
jumps on measurement is that the probabilty of jumps of a spin-1/2 magnetic
moment between two directions separated by an angle ↵ is cos2 ↵/2.
6
Figure 4: A spin-1/2 particle passing through a Stern-Gerlach apparatus ori-
ented in the vertical (Z) direction. Whatever the initial direction of the spin
(this can be prepared from a previous Stern Gerlach apparatus – explained
in the text), it exits oriented either in the +Z direction (i.e., with magnetic
moment +µB in the Z direction) or in the Z direction (i.e., with magnetic
moment µB in the Z direction)
Exercise .5
Consider a large collection of n spin-1/2 particles, all of whose magnetic
moments are oriented along the +Z direction. The particles are sufficiently far
apart so as to not interact among themselves. Show that the net magnetic
moment of all the particles taken together when the magnetic moment of each
particle is measured along the same direction ✓ is nµ cos ✓.
7
bits, the two states of a spin-1/2 along the z axis are depicted as kets |0i and
|1i and represented in the same way as the two states of a bit as represented
in section 2.
1 0
|0i ⌘ , |1i ⌘ (6)
0 1
In the above ket notation, let us represent the general state of a qubit as
| (✓, )i, where is taken as the symbol of a general quantum state following
the historical convention in quantum mechanics, while the variables ✓ and
denote the spherical polar coordinates in the Bloch sphere. Note that the
number of possible states of a qubit (labelled by the two variables ✓ 2 [0, ⇡]
and 2 [0, 2⇡]) is infinite, and thus, for spin states other than those along
the z axis, we have to keep our minds open for the possibility for a more
general representation such as
↵
| (✓, )i = , (7)
where ↵ and can be any number (including the possibility of being complex
numbers) and it is natural to expect them to be some functions of ✓ and .
The spin states along the z axis (Eq.6) are then special cases with ↵ = 1, =
0 and ↵ = 0, = 1 respectively. For these special pair of states, if we take
the row vectors [1 0] and [0 1] corresponding to the states in Eq.(6), and
represent them as h0| and h1| respectively, then, multiplications of di↵erent
row and column vectors in the set {h0|, h1|, |0i, |1i} (with always the row
vector first and the column vector second, so as to yield a single number as
the answer) would give
1 0
h0|0i = [1 0] = 1, h1|1i = [0 1] = 1, (8a)
0 1
0 1
h0|1i = [1 0] = 0, h1|0i = [0 1] = 0. (8b)
1 0
As what one takes as the z axis of the Bloch sphere is completely arbitrary,
then, generalizing the above Eq.(8a), we would like to construct a vector
representation | (✓, )i such that
which is called the normalization condition. If we want the keep the option
open for complex values of ↵ and and yet satisfy Eq.(9) (which implies
h (✓, )| (✓, )i > 0) then one has to define
h (✓, )| = [↵⇤ ⇤
], (10)
8
Exercise .6
2 q 3
2
3
For a ket |⌘i = 4 q 5 find the bra and hence compute h⌘|⌘i. With
i 13
" #
0
p1 0
5
another ket |⌘ i = compute h⌘ |⌘i. Important: Use Eq.(10) and do
i p25
not forget to take the complex conjugates necessary in the definition of the bra.
9
Figure 5: The states of two di↵erent physical systems, the spin-1/2 and the
polarization of a photon on the Bloch sphere.
which can act as qubits. For example, for a polarization state of a photon,
horizontally and vertically polarized states |Hi and |V i can correspond to |0i
and |1i, while 45 and 135 degree polarizations and circular polarizations cor-
respond to other directions as shown in Fig.5. The correspondence between
points on the Bloch sphere and physical states (spin-1/2 and polarization of
light) are shown in Fig.5. Moreover, the correspondence between di↵erent
points and state vectors are shown in Fig.6. This latter figure also shows
important notation |+i and | i states, as well as how nonclasical they really
are in terms of the Schoredinger Cat.
10
θ θ
ψ (θ , φ ) = cos 0 + eiφ sin 1
0 2 2
1
− =
2
(0 −1 )
φ
1 1
2
( 0 −i 1 ) 2
( 0 +i 1 )
1
+ =
2
(0 +1 )
1
Figure 6: The points of the Bloch sphere connected with the state vectors
they represent. Note the important notation of |+i and | i states. The
figure illustrates how non-classical states (states which do not correspond
to our everyday classical experience) reside along the equator of the Bloch
sphere if one identifies the |0i and |1i states with states of classical experience
such as alive and dead states of a cat.
11
as then its square (its application twice) would correspond to a N OT instead
of 1.
were in the last step a phase factor of ei outside the ket has been ignored (as
we mentioned in section 6, a phase factor outside the state of a qubit can be
ignored as long as we are only dealing with that qubit). If you visualize the
state | (⇡ ✓, 2⇡ )i in the Bloch sphere, it will be clear that it corresponds
to the rotation of the state | (✓, )i by an angle ⇡ about the x axis. Thus
X = Rx (⇡), (12)
where the notation Rx (↵) denotes the operator rotating kets in the Bloch
sphere by an angle ↵ about the x axis. Similarly, it is easy to verify that the
operators
0 i
Y = , (13)
i 0
and
1 0
Z= , (14)
0 1
perform rotations by ⇡ about the y and z axes respectively. Note that (verify
yourself by matrix multiplications)
X 2 = XX = 1, Y 2 = Y Y = 1, Z 2 = ZZ = 1, (15)
and
Y = iZX. (16)
12
The identities Eq.(15) will be used very often in constructing circuits for
quantum computation, while because of identity Eq.(16), the operator Y is
used less frequently in quantum information texts (the operators X and Z
are successively applied to implement the operator Y ; the extra phase factor
which will appear is not relevant as long as we are dealing with a single qubit,
and can be taken into consideration in the other cases).
Exercise .7
By applying the matrix form of the operators Y and Z on the vector | (✓, )i
verify that Y = Ry (⇡) and Z = Rz (⇡), were Ry ( ) and Rz ( ) denote rotations
about the y and z axes by angles and respectively.
Exercise .8
Show that HZH = X and HXH = Z (In words, this is called: Hadamard
changes from X basis to Z basis).
12 Qudits
Qubits manifest only two possible fully distinguishable states in any given
measurement (remember the Stern-Gerlach experiment, where for any given
experiment, i.e., for any given orientation of the Stern-Gerlach setup, there
are only two possible outcomes), which we call |0i and |1i. We call these
2-state quantum systems. However, most systems in nature are not 2-state
systems. For example, if you measure the position of an object you obtain
real numbers from an infinite range of possibilities – ideally one should call
it infinite-state quantum system in conformity with calling a qubit a 2-state
quantum system – but in practice, it is called an infinite dimensional quantum
system (here the dimensions have nothing to do with the physical dimensions
occupied by the system, such as the 3 dimensions of our world, but the num-
ber of possible fully distinguishable outcomes a measurement of the system
can give) or simply a continuous variable system. Alternatively, in contrast
to Fig.2, which depicts a double quantum dot charge qubit (each box being
a quantum dot which holds an electron), imagine multiple dots (boxes) in a
row, with an electron having the possibility of being in any one of the dots.
If there are d such dots, teh system is called a quantum d-dimensional system
or a “qudit”.
To appreciate the number of possible states of a qudit, it is advisable to
look at yet another realization of a qubit, namely the state at the slits of a
Young double slit experiment. This is shown in the left hand side of Fig.7.
Here we can recast the notation used for qubits so that we relabel cos 2✓ , sin 2✓
and in terms of arbitrary complex numbers c0 and c1 obeying the condition
|c0 |2 + |c1 |2 = 1 (called the normalization condition). As noted in section 6 a
phase factor outside the state does not matter, so that an arbitrary state of
13
a qubit is represented as
| i = c0 |0i + c1 |1i
14
0 c0
1 c1
0 c0
2 c2
1 c1
. iδ j
. cj = cj e
.
θ θ .
cos = c0 , sin = c1 ,
2 2
iδ j
c j = c j e , φ = δ1 − δ0
d −1 cd−1
Qubits Qudits
15
Lecture 2
Sougato Bose
January 17, 2021
1
1 0 0 1 0 1 1 0 1 0 1 1 1 0 1 0 1 0 0 1 1 0 0 1 1 0 0 0 1 0 0 0
X X X X Z Z Z Z Z Z X X Z Z X Z Z Z X Z X X X Z X X Z Z X Z X Z
− + + − 0 1 1 0 1 0 + − 1 0 − 0 1 0 + 1 − + + 1 − + 0 0 − 0 + 0
X Z Z X X X Z X X Z X Z Z Z X X
+ 0 1 + − − 0 + + 1 + 1 1 0 − +
0 0 1 0 1 1 0 0 0 1 0 1 1 0 1 0
X Z X X Z Z X Z Z X Z Z X Z Z X Z X X X Z X Z Z Z X Z X X Z X Z
− 0 + 0 0 1 + 0 1 − 0 1 + 0 1 + 1 + + − 0 + 0 1 0 + 0 − − 1 + 0
1 0 0 0 0 1 0 0 1 1 0 1 0 0 1 0 1 0 0 1 0 0 0 1 0 0 0 1 1 1 0 0
Eavesd 1 0 0 1 1 0 1 0 1 0 0 1 0 0 1 1 0 0
ropper
detec2
on E K K K K K E
Error
Reconc
0 1 0 0 1 0 1 1 0 0
ilia2on
0 1 0 0 1 0
Privacy
Amplifi 1 0 1
ca2on
The generated states are shown in the third row of Fig.1. Typically in long
distance cryptography, the states will be encoded in photon polarizations as
|0i = |Hi, |1i = |V i, |+i = p12 (|Hi + |V i) = |45i, | i = p12 (|Hi |V i) =
|135i, where H stands for horizontally polarized, V for vertically polarized
photons, while the two other directions mentioned are polarized along the
angles mentioned.
Step 2: Alice transmits the generated quantum states to Bob using some
channel (e.g. for photons it will be an optical fiber). This is shown as the
downward directing blue arrows in Fig.1.
Step 2.5 (Only if Eve is present): Eve does an intercept and resend
attack on the transmitted states. This means that Eve measures a fraction f
of the states that Alice had sent randomly in the x or z basis (Eve’s basis is
depicted in row 5 of Fig.1), encodes the measured state on to a new carrier
(e.g. new photon) (row 6) and sends the state to Bob. Eve generates a bit
2
string key from the measurement outcomes following the same prescription
as Alice’s key generation (row 7). In Fig.1 the fraction measured by Eve is
f = 1/2 – Eve measures every alternate qubit transmitted by Alice. This
is depicted by downward arrows alternating between continuing to Bob and
stopping at Eve, who resends the state (depicted as new arrows emanating
below Eve’s measured states) to Bob. As Eve does not know Alice’s basis
string b, Eve will measure a fraction f /2 in the correct basis and a fraction
f /2 in the wrong basis.
Step 3: Bob measures randomly in the x or z basis according to another
bit-string xxzzz....xzzx of x’s and z’s (row 9). This bit string is di↵erent
from Alice’s bit string b and randomly chosen. So the basis chosen has
1/2 probability to overlap with Bob. Bob generates a bit string (row 11)
according to the states he obtains as outcomes of his measurements (row
10) following the same prescription as Alice’s key generation, i.e., |+i =)
0, | i =) 1, |0i =) 0, |1i =) 1
Step 4: Public comparison of Bases – Alice and Bob publicly compare
their bases (announcing on a public channel: everyone can hear). They keep
only those cases for which their bases matched. The retained cases, where
Alice and Bob’s bases have matched, are depicted in row 12 of Fig.1. The
fraction of retained cases for large strings should be 1/2. However, as we can
see in the figure, due to small number fluctuations, 18 out of the 32 bits are
retained.
Step 5: Evedropper Detection/ Error Estimation – Alice and Bob
now take a small sub-string of their retained bits (this is shown as the purple
coloured section of 6 bits in row 12 of Fig.1) and publicly compare their
outcomes. Note that a fraction of these bits will be known by Eve. These
bits are denoted by K in Fig.1. Moreover, due to Eve’s actions, in a fraction
f /4 of the publicly compared bits Alice and Bob should be errors (see the
discussion of security below). Thus by comparison of a small substring of
their bits, Alice and Bob identify f . If f (quantifying the degree of Eve’s
intervention) is below a desired threshold, then they proceed with further
stages of the protocol. If f is too high, they abort the protocol, and seek
alternate methods to share their secret key. In Fig.1, the bits that have errors
(mismatch between Alice and Bob’s retained bits) are denoted by E in row
13. In the case considered in Fig.1 f = 1/2 and thus errors should be in 1/8
of the bits, but due to statistical fluctuations in the small sample, errors are
actually found in 2 out of 18 bits.
Step 6: Error Reconciliation The aim of this step is to find out the
location of the errors in the private (unannounced) part of Alice’s and Bob’s
retained bit strings (i.e., the 12 bits in the uncoloured part of row 12) and
eliminate them. We know that a fraction f /4 will have errors for large bit
strings. While finding out the errors Alice and Bob have to do it in such a
way that minimal or no extra information is leaked to Eve over what Eve
already knows. One fruitful strategy usually adopted for this is a bisective
search whereby the bit-string is divided to two equal parts down the middle
(in the figure, blocks of 6 each as shown by horizontal brackets in row 13),
their parities (the value of the binary addition of the bits modulo 2) are
3
announced publicly and the first bit of each collection is discarded. The
discarding is done for the purpose of getting rid of Eve’s information about
the parity from the public announcements – as the discarded bit is likely
to be unknown to Eve (for small f ), it gets rid of any information about
the parity of the bits retained after the above discarding. An assumption
in the bisective search is that at most 1 bit contains an error in the blocks
made for parity comparison, which should hold largely true for small f . The
bits remaining after this are shown in row 13. The blocks whose parities
match, are simply continued to the next step (row 14), while those for which
the parities did not match (the right block in Fig.1), a bisective search is
repeated by splitting that block into two further blocks. This is how the
bit labeled by E is identified and eliminated. On the way, because of parity
comparisons, several other bits are also eliminated. The final string after
this (row 14) is now without errors (Alice and Bob have perfect match in
the bits retained). In practice, for realistic larger cases, the bisective search
is followed by other searches to handle cases of two errors per block, and
errors are greatly minimized (to a desired quantifiable extent) rather than
fully eliminated in the end.
Step 7: Privacy Amplification – While the resulting strings (row 14) are
error-free and shared between Alice and Bob, Eve knows a fraction of these
correctly. The target in this step is to compress the shared (now nearly error
free) bit strings to get rid of Eve’s information to the best extent possible (i.e.
enhance the privacy of the key shared between Alice and Bob). In Fig.1, this
amounts to making the information Eve has of the bits marked K irrelevant.
This is done by combining selections of bits and using their parity as the
key (obviously which bits are being combined is publicly announced so that
Alice and Bob have the same final keys). In Fig.1, for example, moving left
to right along row 14, the parities of two successive bits is computed and
used as the final key between Alice and Bob (row 15). Note here that as Eve
knows only one bit within each combined pair, her knowledge is completely
eliminated (a random bit being added binary modulo 2 to the bit she knows).
In parctice for realistic cases, Alice and Bob will reduce Eve’s knowledge of
the key to a certain agreed threshold level.
4 A discussion of security
Let us consider the shared bit strings after the public comparison of bases
by Alice and Bob. Eve has measured a fraction f of the qubits transmitted
by Alice. Of these Eve will measure a fraction 1/2 in the correct basis
and the state sent by Alice will remain intact and be transmitted to Bob.
Thus for these cases, Bob will see no errors in the shared bit string obtained
after the public comparison. Of the rest 1/2 fraction of Eve’s measured
qubits, a fraction 1/2 will probabilistically collapse to the right outcome
as |h0|±i|2 = |h1|±i|2 = 1/2. Thus only 1/4 of those measured by Eve
will have errors, implying a error of f /4 observed by Alice and Bob. In
practice, Eve can attack in ways other than intercept and resend such as
4
by entangling another system with the qubit transmitted or by trying to
copy (clone) it. These also will leave an error trace of similar magnitude.
For example, entangling will increase the mixedness of the states reaching
Bob, with more mixedness for more entanglement Eve attempts to make in
order to get more information, while why Eve cannot clone will be discussed
later in the course. Indeed Eve can use a full quantum computer, which can
in e↵ect do entanglement of larger blocks of qubits with a large number of
ancillary qubits and subsequently measure those ancillary ones, but this will
also leave a trace due to increasing mixedness.
5 Photonic Qubits
Photonic qubits are generally used for QKD as photons can travel long dis-
tances without interacting significantly with their environment so that their
states (say polarization states as an example) are well preserved. However,
photons do not directly interact with each other. This makes their usage in
quantum computation, where interactions can aid in quantum gates very dif-
ficult (although people do accomplish it, albeit less efficiently than would be
possible through direct interactions using clever usage of quantum statistics
and measurements).
5
Lecture 3: Quantum Evolutions and Circuits
Sougato Bose
January 29, 2021
Rz ( ) = e i 2 Z
1
X
= 1+ (i )n Z n (Using the Taylor expansion of ex )
n=1
2
1
that a clockwise rotation by an angle about any axis in space defined by
the unit vector n̂ = ex nx + ey ny + ez nz is given by
Rn̂ ( ) = ei 2 .n̂
= cos 1 + i sin .n̂, (5)
2 2
where = ex X + ey Y + ez Z. In general, of course, any transformation from
0 0
some state | (✓, )i to some other state | (✓ , )i on the Bloch sphere can
be thought of a rotation about some axes.
Exercise .1
Show that det{Rn̂ ( )} = 1, where det{A} means the determinant of the
matrix A.
1 ABC Theorem
Before we move over to consider quantum operations on two qubits, we prove
a theorem. This theorem will be useful to construct Controlled-Unitaries
from a standard two qubit gates such as a CNOT (we will also show “how”
to construct a CNOT shortly). The statement of the theorem is:
Theorem: An arbitrary single qubit unitary operator can be expressed as
U = ei↵ AXBXC in terms of 3 single qubit unitary operators A, B and C
satisfying ABC = 1.
Proof: The proof is by construction. We assume
A = Rz ( )Ry ( ), B = Ry ( )Rz ( ), C = Rz ( + ).
2 2 2 2 2 2
2
By multiplication it is easily verified that ABC = 1. Now,
|00i ! |00i,
|01i ! |01i,
|10i ! |10i,
|11i ! |11i,
or simply by the matrix
2 3
1 0 0 0
6 0 1 0 0 7
CZ = 6
4 0
7
0 1 0 5
0 0 0 1
HIsing = gZA ⌦ ZB
where g > 0 is a coupling strength which has frequency units (h̄g is regarded
as the energy). Verify yourself that
3
0 0
1 0
(a) (b)
1 1
1 0
(c) (d)
Figure 1: Showing four diferent two qubit states and their the reason for their
energy di↵erence. The |01i and |10i configurations have a smaller distance
between charges and hence a higher energy. The zero of the energy is then
redefined as midway between those of the |00i and |01i states.
4
and |11i are energy eignestates with di↵erent energies ±h̄g. Physically it is
easiest to visualize this for two double-dot charge qubits placed one above
another as shown in Fig.1 with the definitions of the |0i and |1i state opposite
for the two qubits. Similarly, for spin qubits, the orientation of one of the
spins along the z axis controls the energy of the second one (It can originate
from the magnetic field in the z direction given out by one spin acting on the
other spin, causing it to align with the former in order to lower its energy,
when the individual energies of the spins due to other external magnetic
fields in the z direction are highly mismatched; Stronger ones can originate
from Coulomb interactions and the quantum indistinguishability of electrons
along with asymmerties posed by di↵erent z directed magnetic fields on the
two spins).
In addition to the above, we will use an extra Hamiltonian applied to
single qubits. This is easiest to visualize for spin qubits, being simply two
magnetic fields of equal strength directed along the z axis. This extra single
quibit Hamiltonian is
H0 = gZA ⌦ 1B + g1A ⌦ ZB
HCZ = H0 + HIsing ,
and to obtain a CZ gate we operate it for a time tCZ = ⇡4 . We can verify that
⇡ ⇡
i ⇡4 ZA ⌦1B i ⇡4 1A ⌦ZB i ⇡4 ZA ⌦ZB
ei 4 e iHCZ tCZ
|00i = ei 4 e e e |00i
i ⇡4 i ⇡4 i ⇡4 i ⇡4
=e e e e |00i = |00i
i ⇡4 iHCZ tCZ i ⇡4 i ⇡4 ZA ⌦1B i ⇡4 1A ⌦ZB i ⇡4 ZA ⌦ZB
e e |01i = e e e e |01i
i ⇡4 i ⇡4 i ⇡4 i ⇡4
=e e e e |01i = |01i
⇡ ⇡
i ⇡4 ZA ⌦1B i ⇡4 1A ⌦ZB i ⇡4 ZA ⌦ZB
ei 4 e iHCZ tCZ
|10i = ei 4 e e e |10i
i ⇡4 i ⇡4 i ⇡4 i ⇡4
=e e e e |10i = |10i
i ⇡4 iHCZ tCZ i ⇡4 i ⇡4 ZA ⌦1B i ⇡4 1A ⌦ZB i ⇡4 ZA ⌦ZB
e e |11i = e e e e |11i
i ⇡4 i ⇡4 i ⇡4 i ⇡4
= e e e e |11i = |11i
i⇡
Thus e iHCZ tCZ =e 4 CZ . When we want to implement CZ in a quantum
computation involving solely two qubits, the phase outside does not matter.
When it is being done on a part of a circuit involving more that 2 qubits, we
⇡
have to remember that an extra ei 4 phase is also present in this implemen-
tation using an Ising Hamiltonian and accordingly do phase adjustments by
local gates on other qubits as needed.
Once the CZ is obtained, we can implement a CNOT following the quan-
tum circuit of Fig.2 exploiting HZH = X.
5
a a
≡
b X b H Z H
H
⎛ 1 0 ⎞
a a ⎜⎜ H iα ⎟⎟
⎝ 0 e ⎠
≡
b U b CH X H
B X H
A
U = eiα AXBXC
Figure 3: Circuit showing the construction of Controlled-U given the ability
to implement CNOT and local unitaries.
6
3 Controlled-U
Armed with CNOT and utilizing the ABC theorem, we will now see how one
can implement a Controlled-U operation. The circuit is given in Fig.3.
To understand why this circuit accomplishes a Controlled-U , we have to
check its operation one by one on each of the computational basis sates. In
Fig.3, a = 0, 1 stand for the bit values of the inputs when computational (z)
basis inputs are taken. Thus,
4 Controlled-Controlled-U s
Given the ability to contruct Controlled Unitaries (single control and single
target), we will now learn how to achieve Controlled-Controlled Unitaries
(two controls and a single target). Consider Fig.4 where we want to imple-
ment Controlled-Controlled-U . In the figure the Controlled Unitaries used
are Controlled-V and Controlled-V † such that V 2 = U (V is a unitary oper-
ation which is the square-root of the unitary U – obviously this is not unique
– any unitary V satisfying V 2 = U will suffice). COnsider the four cases of
di↵erent values for input computational states one by one, and the resulting
operation on qubit c
a = 0, b = 0 =) none of the controlled operations act =) I acts on c.
a = 0, b = 1 =) Only gates controlle by b acts, =) V † V = I acts on c.
a = 1, b = 0 =) Both CNOTs and the last Controlled-V acts, =)
V V † = I acts on c.
a = 1, b = 1 =) All Controlled gates act, =) V V = V 2 = U acts on c.
Note: One of the most important Controlled-Controlled-U gates is the
TOFFOLI or Controlled-Controlled-NOT. It is the fundamental gate for Re-
versible Classical Computation (all classical circuits can be constructed from
the TOFFOLI). Moreover, TOFFOLI and Hadamard together constitute a
complete set of gates for universal quantum computation. To accomplish a
TOFFOLI
p according to the methodology described here, one has to choose
V = X.
7
a
b X X
c VH VH+ VH
8
a
b
c
0 X X 0
0
X X 0
d U
H
9
. Then 2 3
1 d0 g
V01 U = 4 0 e0 h 5 .
c f i
Similarly applying another unitary operation V02 (similarly chosen for the
levels |0i and |2i) we can get
2 3
1 d0 g 0
V02 V01 U = 4 0 e0 h 5 .
0 f i0
Now a last unitary V12 acting on the levels |1i and |2i (essentially the inverse
of the 2 ⇥ 2 matrix in the |1i,—2i subspace) can bring it to the form
Thus U = V01† V02† V12† is the implementation of the arbitrary qutrit unitary in
terms of three 2 level unitaries.
|000i $ |100i
|100i $ |110i
Now a unitary operation U is done on the third qubit and the above states
are repeated in the reverse order to take the |110i level back to |000i level.
This whole procedure is thus equivalent to performing the 2-level unitary U
in the {|000i, |111i} subspace.
10
7 Universal Quantum Computation
Universal quantum computation means being able to accomplish arbitrary
SU (2n ) on n qubits. From the above discussions it is clear that we can do
that if we could do arbitrary local unitaries and multi-control versions of
TOFFOLI for larger numbers of qubits with the control value being modi-
fied to 0 if needed. These, are, in turn, accomplished by CNOTs and local
unitaries as shown in earlier parts of this lecture. Thus
(we do not prove here why this discrete set suffices to achieve any local
unitary operation).
Another interesting universality avenue is:
Universal quantum computation is possible with a TOFFOLI and Hadamard
H.
We do not prove the above here either.
11
Entangled States: Nonlocality & Dense Coding
Sougato Bose
February 5, 2021
1 Entangled States
Consider a generic ket representing a certain state to two quantum systems
A and B, given by X
| iAB = cjk |jiA ⌦ |kiB .
jk
We will often refer to | iAB as the joint or combined state of two systems as
it represents the total state of both the systems. It is not guaranteed that
we can write
| iAB = | iA ⌦ | iB .
In some special cases we can; for example
(1) 1
| iAB = p (|0iA ⌦ |0iB + |0iA ⌦ |1iB ) = |0iA ⌦ |+iB . (1)
2
In these cases, we call the above state factorizable or a product state. The
factorization works by taking out, as a common factor, a ket |⇠i↵ , from the
combined state of two systems where, in all the terms, both the ket label ⇠
and the system label (subscript) ↵ (in the above example, ↵ = A, B) must
be the same. In the above example Eq.(1), |0iA is common in both the first
and second term of the sum (superposition) of states in | (1) i, and hence it is
factored out in order for system B to be in the state p12 (|0iB + |1iB ) = |+iB .
Consider, now, instead, the example,
(2) 1
| iAB = p (|0iA ⌦ |0iB + |1iA ⌦ |1iB ). (2)
2
You cannot take out either |0iA or |1iA as a common factor from the two
terms as they are not the same state. You cannot take |0iB or |1iB as a
common factor either because of exactly the same reason. Thus | (2) i is
a non-factorizable state. It is, in fact, a canonical example of an entangled
state. As long as we are reperesenting the combined (or joint) state of the two
systems by a ket (we can only do that when they are pure states), whenever
| iAB 6= | iA ⌦ | iB , (3)
we call the state | iAB an entangled state. This definition has to be slightly
generalized when we consider mixed states of A and B as then we cannot use
1
a single ket to represent their joint state – this will be discussed in a later
lecture. If we have multiple systems 1, 2, 3, ..., n whose state is represented
by a ket
| i12...n 6= | 1 i1 ⌦ | 2 i2 ⌦ ... ⌦ | 3 i3 , (4)
we say it is an entangled state.
Exercise .1
Show that
(3) 1
| i = (|00iAB + |01iAB + |10iAB + |11iAB )
2
is a product state. Show that
(4) (3)
| i = CZAB | i
is an entangled state.
2 Quantum Nonlocaity
Entangled states demonstrate quantum nonlocaity when they are shared
among distant observers who do measurements on them. For example com-
sider the case when two qubits in the state | (2) iAB are shared between two
distant parties Alice and Bob; A being held by Alice and B being held by
Bob. Then quantum mechanics exhibits a peculiar feature that the corre-
lations obtained from measurements on | (2) iAB do not follow the locality
assumption: The choice (in this case the choice of one among two possible
measurement bases) made by Alice cannot a↵ect the measurement result
obtained by another observer Bob when they are space-like separated (out-
side each other’s light cone so that no physical signal can propagate between
them). Note that here that the locality assumption is merely demanding
that it is the choice of basis of one observer which should not influence the
measurement outcome of another observer. The choice of basis by Alice can-
not, in fact, provide any bits of information to the distant observer Bob. Bob
gets the same probability distribution of outcomes for all di↵erent choices of
Alice’s basis, and hence cannot infer what basis Alice had chosen (this will
be shown exactly in terms of the density matrix in a later lecture) and there-
fore cannot learn any bits of information from Alice. Thus there is no faster
2
X Y
A
m x myA
Y
myB Y
myC
X X
mxB mxC
Figure 1: The setting for the demonstration of quantum nonlocality by the
GHZ state. Each observer measures her/his qubits in the X or Y basis with
outcomes labelled by mA A
x , my and similarly for B and C.
3
value (either +1 or -1) is independent of whether Bob and Charlie measure
XB YC or YB XC . We will show that this cannot be true by the following
contradiction.
Note that (verify by calculation remembering that XA , YB and YC act
on the first, second and third qubits of the state respectively, with X|0i =
|1i , X|1i = |0i , Y |0i = i |1i , Y |0i = i |1i)
XA YB YC |GHZiABC = 1|GHZiABC =) mA B C
x my my = 1 (5a)
YA XB YC |GHZiABC = 1|GHZiABC =) mA B C
y mx my = 1 (5b)
YA YB XC |GHZiABC = 1|GHZiABC =) mA B C
y my mx = 1 (5c)
The locality assumption has been used in the rightmost equalities of Eqs.(5a)-
(5c) in the sense that the same mAy (whether it be +1 or -1) occurs in Eqs.(5b)
and (5c) and similarly for the other outcome values. Multiplying the right-
most equalities of Eqs.(5a)-(5c) and using the fact that (mA 2 B 2
y ) = (my ) =
(mC 2
y ) = 1 we get
mA B C
x mx mx = 1. (6)
On the other hand the outcomes mA B C
x , mx and mx should be obtained when
X basis measurements are performed on all 3 qubits (again by the local-
ity assumption that mA
x does not depend on whether Bob and Charlie are
measuring in the X or Y bases, and similarly for the other combinations).
However, quantum mechanics gives (check by direct application of the oper-
ator XA XB XC on the GHZ state)
XA XB XC |GHZiABC = +1|GHZiABC =) mA B C
x mx mx = +1. (7)
The direct contradiction of Eq.(6) and Eq.(7) implies that the locality as-
sumption must have been wrong (i.e., the value mA x is indeed influenced by
the context of the measurement – whether Bob and Charlie choose to measure
in the X or the Y basis – thus there is indeed a spooky action at a distance).
3 Bell-States
There are 4 canonical entangled states of 2 qubits which are called the Bell
states (in honour of John Bell who used one of these class of states first to
demonstrate the nonlocal nature of quantum correlations):
+
↵ 1
AB
= p (|0iA |1iB + |1iA |0iB )
2
↵ 1
AB
= p (|0iA |1iB |1iA |0iB )
2
+
↵ 1
AB
= p (|0iA |0iB + |1iA |1iB )
2
↵ 1
AB
= p (|0iA |0iB |1iA |1iB )
2
4
In fact, the above states are in the class of the most entangled states pos-
sible for two qubits – we will discuss that when we discuss the quantification
of entanglement in another lecture. These states form a complete orthonor-
mal basis for two qubits (please verify for yourself by directly obtaining the
relevant inner products)
± ± + +
h | i=h | i=h | i = 0,
+ + + +
h | i=h | i=h | i=h | i = 1.
Thus a projective measurement on two qubits can be performed on this basis
to obtain one of the Bell states – this is called the Bell basis measurement
or Bell State Measurement (BSM). It is a complete basis for measurement
as there are only 4 possible orthogonal states for two qubits. In a later
lecture we will discuss how Bell states can be generated and how a BSM
can be performed both via a quantum circuit (possible for qubits between
which an interaction can be switched on to enact a CNOT gate) and how
they have been done in practice for noninteracting photonic qubits exploiting
quantum indistinguishability. For the time being, however, we will proceed
by assuming a BSM to be possible.
An interesting and important property of the Bell-States are their in-
terconvertability to each other simply through the action on one of the two
qubits. Such an operation, which acts only on an individual qubit is called
a local operation. For example, XA ⌦ 1B is regarded as a local operation as
the X operator only acts on qubit A and no operation acts on qubit B (i.e.,
it is acted on by the Identity operation).
Verify for yourself (by calculation remembering that operators XA , ZA
and IB act on their respective qubits with X|0i = |1i , X|1i = |0i , Z|0i =
|0i , Z|1i = |1i)
↵ ↵
XA ⌦ 1B + AB = +
AB
(8a)
↵ ↵
ZA ⌦ 1B + AB = (8b)
+
↵ ↵AB
ZA XA ⌦ 1B AB
= AB
(8c)
Exercise .2
Show that the set of states defined by
↵
GHZ srt = ZAs XAr ⌦ XBt ⌦ 1C |GHZiABC
5
Ψ+
AB
(a)
α ⇒ IA ⊗ IB
β ⇒ XA ⊗ IB
γ ⇒ ZA ⊗ IB
δ ⇒ ZA XA ⊗ IB (b)
BSM
Ψ+ ⇒ α
Ψ− ⇒ β
Φ+ ⇒ γ
(c)
Φ− ⇒ δ
Figure 2: Quantum Dense Coding protocol showing three of the steps of the
protocol.
Step 1: Alice and Bob share many copies of the entangled state | + iAB of
two qubits through a quantum channel (a channel which can transmit quan-
tum states, in this case the state of a qubit). This is ideally done overnight
(or more generally in a non-rush hour time period) so that we do not really
need to “count” these transmissions as a significant load to the channel –
the channel is mostly unused at these times. Note no information is actually
transmitted at this stage from Alice to Bob – only the correlations entailed by
6
the state | + iAB . In fact, suppose that the information Alice will eventually
send to Bob has not even materialized (e.g., it could be the news emerging
from an event occuring the next day). The situation after the completion of
this step, depicting only one of the copies of | + iAB is shown as Fig.2(a).
We will describe the protocol using this single copy (with the understanding
that each copy will be used in exactly the same way)
Step 2: Alice now has information to send to Bob. She can now encode
this information on her qubit at the rate of 2 bits/qubit – thus dense coding
(twice the capacity as is possible for a single qubit). She performs one of
four possible operations on her qubit. Communicating 2 bits correspond
to communicating four letters – say we call these ↵, , and . Now Alice
performs the following protocol to encode her letter to the qubit she has (this
is shown in Fig.2(b)):
↵ =) 1A (no action) on qubit A
↵ ↵
=) State of A and B = 1A ⌦ 1B + AB = +
AB
=) XA on qubit A
+
↵ +
↵
=) State of A and B = XA ⌦ 1B AB
= AB
=) ZA on qubit A
+
↵ ↵
=) State of A and B = ZA ⌦ 1B AB
= AB
=) ZA XA on qubit A
+
↵ ↵
=) State of A and B = ZA XA ⌦ 1B AB
= AB
+
In the above the transformation of | iAB to the other states by Alice’s local
operation follows from Eqs.(8).
Step 3: Alice transmits her qubit to Bob. She intends Bob to learn the 2
bits of information she encoded from this. Of course, she encodes several such
qubits, one for each copy of | + iAB shared with Bob, with the intention that
each will communicate 2 bits of information (2 bits per transmitted qubit).
Note that these transmissions are occuring during the rush hour, so we do
count these transmissions as a load on the channel – however, now only one
qubit per two bits is being transmitted through the quantum channel (instead
of one qubit per bit).
Step 4: Bob receives Alice’s qubit and measures both qubit A and B
in the Bell basis (Bell State Measurement BSM, as schematically shown in
Fig.2(c). He gets four possible outcomes, from which he infers the 4 possible
letters sent by Alice as follows:
↵
Bob obtains + AB =) Alice sent = ↵,
↵
Bob obtains + AB =) Alice sent = ,
↵
Bob obtains =) Alice sent = ,
↵AB
Bob obtains AB
=) Alice sent = .
Thus Bob obtains 2 bits of information per qubit transmitted by Alice after
encoding.
7
Exercise .3
Check that no dense coding will happen with the state |01iAB shared between
Alice and Bob (i.e., Alice will communicate at most 1 bit per qubit).
The above exercise illustrates the importance of sharing an entangled
state in the dense coding protocol.
8
Quantum Teleportation
Sougato Bose
February 9, 2021
After dense coding, we will now study the most celebrated application of
entanglement, namely Quantum teleportation. Quantum teleportation is the
act of sending a quantum state accross a distance using shared entanglement
in a way that the original quantum state is destroyed, and a copy reappears
at a distance. Thus it is a very di↵erent strategy from the usual way you
send a “scanned” letter to a distance. You essentially make a copy of the
information which is sent over the internet, while the original remains with
you. Hwoever, this strategy does not work for quantum states – you cannot
make a copy – this is also the reason why QKD is secure – Eve cannot make an
exact copy of the quantum state transmitted by Alice to Bob. Thus before
studying teleportation, we should study the quantum no-cloning theorem,
which proves that you cannot construct a machine which makes a copy of an
arbitrary quantum state.
1
quantum state, it should also clone | i + | i – linear combination of the two
states cloned above in Eqs. (1a) and (1b), so that we must also have
In writing the above equation, we have ignored the issue of normalizing the
state | i + | i as that is not relevant to the logic of the proof. However, from
Eqs.(1a) and (1b), and the linearity of L it follows that
Clearly, the states on the right hand side of Eqs.(2) and (3) are not the
same as long as | i and | i are not the same (i.e. | h | i| = 6 1). As Eq.(3)
follows from linearity of the evolution operator in the theory, it has to be
correct. This automatically implies that Eqs.(2) is incorrect, and thereby
the | i + | i cannot be cloned. Thus arbitrary states (i.e., any random state
of your choice) cannot be cloned by a machine – or in other words, a universal
cloning machine which clones quantum states cannot exist.
and thereby the normalized state of the rest of the particles being
1 ˜ DEF .
| (m)iDEF = p | (m)i (6)
Pm
Note that although the right hand side of Eq.(4) superficially looks like an
inner product, it is still not so as hm|ABC only acts on part of the state; it
is thus still a ket (depicting the state of the remaining 3 qubits as in the
left hand side of Eq.(4)). Eqs.(4)-(6) will be used extensively in our study of
teleportation. However, they are essentially the same as the more standard
2
Ψ+
AB
ψ C
=α 0 C
+β 1 C (a)
ψ B
BSM
ZB ψ B
XB ψ B
+ − + −
Ψ ,Ψ ,Φ ,Φ XB ZB ψ
(b)
CA CA CA CA
B
2 bits
ψ B , ZB ψ B , XB ψ B , XB ZB ψ B
⇓ IB ⇓ ZB ⇓ XB ⇓ ZB XB
ψ B
ψ B
ψ B
ψ B
(c)
˜ DEF ,
⇧m | iABCDEF = |miABC | (m)i
˜ DEF is the same as that in Eq.(4). Similarly, in terms of the
where | (m)i
standard projection operator ⇧m , simply by normalizing the above unnormal-
ized state, the appropriately normalized state of the whole system is (verify
by yourself)
⇧m | iABCDEF
|miABC | (m)iDEF = p .
Tr(⇧m | ih |ABCDEF ⇧m )
3 Quantum Teleportation
We describe quantum teleportation in the following steps:
Step 1: Alice shares an entangled state | + iAB with Bob, with Alice holding
3
qubit A and Bob holding qubit B.
Step 2: A qubit in an arbitrary and unknown state | iC = ↵ |0iC + |1iC
encoded in a qubit C is brought to Alice. Her task is to send this state over
a distance to Bob using measurements and classical communication (com-
munication of bits of information, but no qubits). The scenario combining
Steps 1 and 2 are shown in Fig.1(a). The joint state of the qubits C, A and
B is thus ↵
| iCAB = | iC + AB .
Step 3: Alice now conducts a Bell state measurement (BSM) on the
qubits C and A. This results in 4 di↵erent states of Bob’s qubit B corre-
sponding to the 4 possible outcomes (| + iCA , | iCA , | + iCA , | iCA ) of
Alice’s BSM as is depucted in Fig.1(b). We calculate the states of qubit B
for the various outcomes of the BSM, and the probabilities of these outcomes,
by using Eqs.(4)-(6) as follows:
↵ E
Outcome +
CA
=) Unnormalized state of qubit B = ˜(1)
⌦ + 1 1
= CA
| iCAB = p (h01|CA + h10|CA )(↵ |0iC + |1iC ) p (|01iAB + |10iAB )
2 2
1 1
= (↵ |0iB + |1iB ) =) ProbabilityP| + i = h ˜(1) | ˜(1) i =
2 ↵ 4
=) Normalized state of qubit B = (1) B = ↵ |0iB + |1iB = | iB .
(7a)
↵ 1
Outcome CA
=) Probability = , Normalized state of qubit B
(2)
↵4
= B
= ↵ |0iB |1iB = ZB | iB .
(7b)
↵ 1
Outcome + CA =) Probability = , Normalized state of qubit B
(3)
↵ 4
= B
= ↵ |1iB + |0iB = XB | iB .
(7c)
↵ 1
Outcome CA
=) Probability = , Normalized state of qubit B
↵ 4
(3)
= B
= ↵ |1iB |0iB = XB ZB | iB .
(7d)
In the above, we have only explicitly shown the calculations for Eq.(7a), while
identical calculations for the projections to the other 3 Bell states would yield
the results of Eqs(7b)-Eqs(7d).
Step 4: Alice communicates her BSM outcome to Bob. This is equiva-
lent to communicating 2 bits of information (one possibility of 4 outcomes).
4
follows:
+
↵
Outcome =) Bob performsIB (nothing)
CA
↵
=) Bob’s state = IB (1) B = IB | iB = | iB . (8a)
↵
Outcome CA
=) Bob performsZB
(2)
↵
=) Bob’s state = ZB B
= Z B ZB | i B = | i B . (8b)
↵
Outcome + CA =) Bob performsXB
↵
=) Bob’s state = XB (3) B = XB XB | iB = | iB . (8c)
↵
Outcome + CA =) Bob performsZB XB
↵
=) Bob’s state = ZB XB (4) B = ZB XB XB ZB | iB = | iB . (8d)
Note that in each case, Eqs.(8a)-(8d), the initial state of C has exactly ap-
peared on the qubit B held by Bob. This concludes the process of telepor-
tation.
Exercise .1
Consider the entangled state of two qubits p12 (|0iA |+iB + i |1iA | iB ) being
used to teleport a state ↵ |0iC + |1iC . Find out the operations Bob should
perform in order to complete a teleportation process if Alice is using a standard
Bell-State measurement procedure.
5
where ebit is the unit with which we quantify the entanglement of 2 qubits
(again, this will become clearer – will be defined – in the next lecture), with
| + iAB possesing an ebit of entanglement. Note that the above equality is
not for changing sides – if you took qubit to the left or bits to the right, it
would not be a valid protocol. In a similar vein the dense coding protocol
can be written as
1ebit + 1qubit = 2bits.
In both the teleportation and dense coding protocols, what is common is
that they both require a shared | + iAB , i.e., 1 ebit.
6
Bell State Measurements and Teleportation
Applications
Sougato Bose
February 10, 2021
In this lecture we are going to complete all the requirements for the
entanglement protocols we have presented, namely show how to do the Bell
State measurements, and then go on to present some of the applications
of teleportation and teleportation like operations in quantum information
processing.
+
↵ CN OTAB 1 HA
AB
! p (|0iA + |1iA ) |0iB ! |0iA |)iB .
2
+
Thus detecting |0iA |0iB amounts to detecting the Bell state | iAB .
Exercise .1
1
H 0 A
Φ+
AB
X 0 B
Figure 1: The quantum circuit for Bell State Measurement is shown with the
Bell state | + iAB as the input, which evolves to |00iAB . When the qubits
are detected in the state |00iAB , we conclude that the Bell state put into the
circuit was | + iAB . Other Bell state inputs give other computational basis
(z-basis) outputs.
Find out the outputs of the circuit of Fig.1 for the inputs | + iAB , | iAB
and | iAB and hence justify how measuring the two qubits at the output of
the circuit in the computational basis (z-basis) is equivalent to a BSM.
Exercise .2
Show that the circuit of Fig.1 run in reverse generates the 4 Bell states from
input states in the computational basis.
2
where |ni depicts a state with n bosons of wave-vector k and polarization .
A beam splitter has momentum/wave-vector states as inputs – particles
impinge from one of two given directions A and B (also called channels or
ports) and exits through one of two given ports C and D. Thus A, B, C
and D denote the momentum/wave-vector directions k. The BS reflects the
incident particle with some amplitude and transmits it with the rest of the
amplitude. Thus it essentially does a unitary transformation on the creation
operators of the modes, thus:
BS a†D, + ia†C,
a†A, ! p (2a)
2
† †
BS aC, + iaD,
a†B, ! p . (2b)
2
Essentially, in the above, the reflected channel (e.g. C is the channel obtained
by reflection of A) gets a i phase, while the transmitted channel continues
without phase change.
Let us first consider the case depicted in Fig.2(a) with two bosons (e.g.
photons) incident from opposite ports of the BS with the same polarization
H (horizontal). This state thus has the following evolution
Exercise .3
Show that in the analogous case of two fermions in the same spin state
incident at a BS they anti-bunch (leave through opposite output ports of the BS
as shown in Fig.2(b).
Exercise .4
Show that in the analogous case of two bosons (e.g. photons) incident at a
BS in the state | + iAB = p12 (|HiA |V iB + |V iA |HiB ) bunch at a BS, while
| iAB = p12 (|HiA |V iB |V iA |HiB ) antibunches at a BS.
The above exercise provides
Fact 2: The photonic version of the Bell state | + iAB gives bunching at the
outputs, and
3
a +A,H
A C
BS
B D
a +B,H
(a)
a +A,↑
A C
BS
B D
a +B,↑
(b)
Figure 2: The bunching of two bosons (a) and the anti-bunching of
two fermions (b) incident on a beam-splitter (BS) from opposite ports
(momenta/wave-vector channels) A and B. The internal states of both
bosons are assumed to be the same in (a), i.e., they are both in the hor-
izontally (H) polarized state, while the internal states of both fermions are
also both assumed to be the same in (b), i.e., they are both in the spin state
".
4
Fact 3: The photonic version of the Bell state | iAB gives antibunching at
the outputs.
Exercise .5
Using facts 1, 2 and 3, above, and the action of a PBS as described above,
argue that
(a) for the Bell state | + iAB photon detectors D1 and D2 or D3 and D4 will
click,
(b) for the Bell state | iAB photon detectors D1 and D3 or D2 and D4 will
click,
(c) for the Bell states | ± iAB photon detector D1 or D2 or D3 and D4 will click
(both photons will be detected by the same detector).
From the above exercise, it is clear that the apparatus of Fig.3 can detect,
even without any interactions between the photons, just by their individual
evolutions at the BS and aspects of their indistinguishable behaviour as en-
coded through the bosonic creation operators a†k, (Eq.(1a)), 3 alternatives:
| + iAB or | iAB or | ± iAB . This still enables dense coding, albeit with a
slightly reduce capacity, as 3 letters (instead of the 4 possible for interacting
qubits) can be transmitted by transmitting only a single qubit. In context
of teleportation, we can similarly teleport perfectly for two of the outcomes
| + iAB or | iAB and declare the process as a teleportation to a state with
partial overlap with the transmitted state for the other set of outcomes.
2 Entanglement Swapping
We will use the methodology of “Measurements on a subsystem of a quantum
system” from Lecture-4-Part B here. Consider the situation depicted in Fig.4.
Particles A and B are in an entanged state | + iAB , while particles C and
D are in an entanged state | + iCD . A and C may be light years apart, yet
when B and D are brought together and a BSM is performed on them, with
the outcome | + iBD , then A and C get projected to the state
⌦ + +
↵ +
↵
BD AB CD
1 1 1
= p (h01|BD + h10|BD ) p (|01iAB + |10iAB ) p (|01iCD + |10iCD )
2 2 2
1 E
= p (|01iAC + |10iAC ) = ˜+
2 2 AC
↵ D 1
= + AC (normalized state) with probability = ˜+ ˜+ i = .
4
5
D1
D2
PBS1
± A C
Φ
AB
OR BS
B D PBS2
Ψ±
AB
D3
D4
Figure 3: The apparatus for performing partial BSM for photons. From
“which” of the combination of detectors D1 , D2 , D3 and D4 have clicked, it
is possible to fully discriminate between incident states | + iAB , | iAB and
both of them with | ± iAB , although it is not possible to distinguish be-
tween | + iAB and | iAB (thus 3 possible alternatives can be distinguished:
| + iAB or | iAB or | ± iAB ).
6
Figure 4: The schematic of entanglement swapping.
Exercise .6
Consider the situation depicted in Fig.5, where the 3, 4 and 5 particle entan-
gled states – shown as particles connected by lines are p12 (|000i+|111i), p12 (|0000i+
|1111i and p12 (|0000i + |1111i) respectively. The dashed line connects the pair
of particles on whom a BSM is conducted. Prove that the entanglement of the
particles evolves from the situation depicted in the left hand side of the figure to
the right hand side of the figure as a result of the BSM.
Exercise .7
Generalize the quantum circuit for BSM to a 3 qubit circuit that performs
a GHZ basis measurement given by a projection to any one of the 8 states
|GHZ srt i = ZAs XAr ⌦ XBt ⌦ 1C |GHZiABC , with s, r, t = 0, 1.
Exercise .8
7
Figure 5: Multiparticle entanglement swapping.
8
munination between the parties possessing them (LOCC – local operations
and classical communication). What is a weakly and what is a perfectly en-
tangled pair of particles is the subject of the next part of the lecture. Here
we only emphasize that even if one wants to share an ideal Bell state between
two distant parties, one or more particles have to traverse a quantum channel
in which they loose their high entanglement. Thus the two distant parties
should share a large number n of copies of degraded entanglement to obtain,
form it, a fewer number k n of near perfect entangled states | + i. These
can then be used to connect distant quantum registers via teleportation.
9
Quantification of Pure State Entanglement:
von Neumann Entropy
Sougato Bose
February 12, 2021
1 Shannon Entropy
Consider mutually exclusive random events j which can occur with probabili-
ties Pj . Suppose someone (say, Bob) does not know which event has occured.
What is the minimum number of bits of information (essentially the num-
ber of yes/no questions) on average required to specify to this person the
event that has actually occurred? The answer is given by the Shannon en-
tropy. Before Bob learns which event has occurred, he has some uncertainty
in knowledge. Thinking about the same question in reverse, the amount of
uncertainty of Bob before he learns exactly which event has occurred is also
the same number of bits (the Shannon entropy), as learning the answer re-
moves his uncertainty. In information theory textbooks such as Cover and
Thomas, they prove that (by using a result called Kraft’s inequality) that for
the most efficient average length of codes, the number of bits used to convey
a specific outcome scales as nj = log2 (1/Pj ) – rarer the event we use more
bits. But you can also understand it heuristically (not rigorously!) from
common sense thinking as an answer to a ”yes/no” game where you have to
arrive at the right event after asking the least number of questions. It makes
sense to first ask whether the most probable one (say, there is an event with
probability ⇠ 1/2) was realized. If it was, yes (yes could stand for the bit
0), we have just used one bit to convey the information! If no (bit 1), we ask
1
whether the next most probable incident (say, something with probability
⇠ 1/4) was realized or not. If yes, then with the two bit string 10, we have
communicated that event, and so on (next probable event with probability
⇠ 1/8 would be communicated with the bit string 110, and so on). In the
above, you see that for a event with probability ⇠ 1/2nj , we are using nj
bits. Such codes are also called intantaneous or prefix codes – no code is
the prefix of another, so that there is no confusion when you get a string of
events (whenever there is a 0, we know we are ending the code of a given
event and starting a new one). Thus the average number of bits to convey
mutually exclusive random events j which can occur with probabilities Pj is
X X X
hnj i = Pj n j = Pj log2 (1/Pj ) = Pj log2 Pj
j j j
Thus if we define
X
⇢= Pj | ji h j| , (1a)
j
2
non-zero, the system is not described a unique ket, and thereby it is called a
mixed state. When a system is in a state which is exactly represented by a
single ket, it is called a pure state.
The density matrix has the P following properties (in the expressions below
{|mi} forms a complete basis, m |mi hm| = 1)
X X
Tr(⇢) = hm| ⇢ |mi = Pm = 1, (2a)
m m
(using def. of trace,& total prob. = 1)
⇢ = ⇢† (2b)
(Hermitian; checked from def. of ⇢),
hm| ⇢ |mi = Pm > 0 (2c)
(as it is a probability).In short depicted as⇢ > 0(a positive operator)
⇢2 = | ji h j| | ji h j| =| ji h j| = ⇢ =) Tr(⇢2 ) = 1. (3)
where the states {| j i} are not necessarily orthonormal. This is unlike the
classical probability distribution Pj of events j discussed in the section on
Shannon entropy because, in the generic case of non-orthonormal {| j i} the
events occuring (di↵erent kets | j i) are not fully distinguishable as the classi-
cal events j. So in some sense
P the state has less uncertainty than quantified
by the Shannon entropy j Pj log2 Pj . Conversely you will need to learn
a lower number of bits of information to become certain of the state. To
treat the density matrix as a classical distribution we need its expression in
a diagonal form in an orthonormal basis with positive eigenvalues so that
we can treat each state of that basis as equivalent to a fully distinguishable
classical event and the eigenvalues as the probabilities of those events. This
is possible as the density matrix is a Hermitian operator (it is diagonaliz-
able with orthonormal eigenstates | j i), while the positivity of ⇢ ensures
j = h j | ⇢ | j i > 0 can be interpreted as probabilities. Thus it is always
possible to write X
⇢= j | ji h j| .
j
3
Thus the number of bits of uncertainty in ⇢ (or, given ⇢, the number of bits
you need to learn in order to become certain of the state) is given by the
Shannon entropy X
j log2 j
j
of the density matrix expressed in its eigenbasis. However, note that the
Shannon entropy in this particularly basis is equivalent to
X
S(⇢) = Tr(⇢ log2 ⇢) = j log2 j (5)
j
Exercise .1
0
Find the von Neumann entropy of the state ⇢ = 12 |0i h0| + 12 |+i h+| and
show that it is lower than (a) the Shannon entropy of a probability distribution
1
2
for event 1 and 12 for event 2, (b) the von Neumann entropy of the state
00
⇢ = 12 |0i h0| + 12 |1i h1|.
Aside the interpretation given above, the von Neumann entropy also has
a few other interpretations:
(a) If we try to communicate with quantum letters | j i with them being used
with probabilities Pj (think of English – you use e much more frequently than
z – we are generalizing this to a quantum alphabet with letters |ei and |zi not
necessarily orthogonal), then the average number of bits we can communicate
per letter (the classical capacity of the quantum alphabet in terms of numbers
of bits) is given by
X
C = S(⇢) with ⇢ = Pj | j i h j | .
j
While exact proof of the above is beyond the scope of this lecture, the in-
tuition is very clear from the arguments above that S(⇢) bits on average
per qubit is the information possible to be held in ⇢. This is known as the
Holevo-Schumacher-Westmoreland theorem for pure state letters. For mixed
state letters ⇢i occuring with probabilities Pi it can also be generalized to
X X
C = S(⇢) Pi S(⇢i ) where ⇢ = P j ⇢j .
i j
(b) The von Neumann entropy S(⇢) also quantifies the compressibility of
a quantum state (called Schumacher compression after Ben Schumacher, who
also coined the term “qubit”). If there are n qubits, each in a mixed state
⇢ , then only a smaller number m = nS(⇢) of qubits need to be faithfully
transmitted through a channel in order to reconstruct the joint mixed state
of the n qubits (i.e, ⇢⌦n ) faithfully at a distance. Again the proof is beyond
the scope of this lecture.
4
(c) In this lecture, we will show (with a proof) the most classic application
of the von Neumann entropy – that it is a quantification of entanglement of
two systems in a pure state. But before that, we need to introduce the
reduced density matrix.
Using a complete basis {|kiB } of states for system B we insert 1A = |ki hk|A
in the above expression to obtain
X X
Pm,A = c⇤i cj h⇠i | mihm |⇠j i h i | kihk| j i
ij k
X X
= c⇤i cj hm |⇠j i h⇠i | mi hk| j i h i | ki
ij k
X
= hm|A ( hk|B | iAB h |AB |kiB )|miA
k
5
then we can obtain all relevant probabilities of outcomes on system A from
it. The above definition has been motivated by an initial pure state of the
two systems described by ⇢AB = | i h |AB , but it is easy to verify that this
definition also holds for mixed states described by ⇢AB .
We will, from now on, drop the superscript “red” for reduced in the den-
sity matrix, and simply use ⇢A to denote the reduced density matrix of a
system A. The deep reason for this is that when we have any given density
matrix, it is always possible to regard it as being part of a larger collection of
systems – whether it is entangled with them or not. So the term “reduced”
looses any intrinsic meaning. However, the method of computing the reduced
density matrix remains as outlined in Eq.(7) even if we are just refer to it as
⇢A .
Exercise .2
Using the above observation immediately write down (without calculations)
the reduced density matrices
q ⇢A and ⇢B for q
a combined state of two systems A
0↵
and B given by AB
= 23 |0iA |1iB + i 13 |1iA |0iB .
Exercise .3
Find out the reduced density matrices ⇢A and ⇢B for a combined state of two
systems A and B given by | iAB = p12 |0iA |0iB + p12 |1iA |+iB . (You have to
do a little work out to find ⇢A , while ⇢B can be immediately written down from
the observation above.)
5 Schmidt Decomposition
We now state a very powerful theorem (without proof) which is a key result
in the study of entanglement of two systems whose joint state is a pure state
6
(described by a single ket). This theorem is equivalent to a result from Linear
Algebra called the Singular Value Decomposition. This result states that the
combined state of two systems A and B can always be written as
Xp 0
E
| iAB = j | i
j A j . (8)
B
j
0↵
where each of {| j i} and { j } form an orthonormal basis (h j | k i =
⌦ 0 0
j k i = jk ). The sum over j extends to the dimension of the Hilbert
space of the smaller of the systems A and B – for example, for a qubit-qutrit
entangled state j would range over only 2 values 0 and 1. You might think
that Eq.(8) is a trivial result as one can always appropriately rewrite any two
particle state, say,
1 1 1 1 1
p |0iA |0iB + |1iA |0iB + |1iA |0iB = p |0iA |0iB + p |0iA |+iB .
2 2 2 2 2
However, note that in the above the set of states {|0iB , |+iB } is not an or-
thonormal basis, and thereby, Schmidt decomposition is a non-trivial result.
It guarantees that a decomposition in terms of orthonormal bases exists as
given by Eq.(8). The Schmidt decomposition and our observation in the
preceding section on reduced density matrices immediately imply that the
reduced density matrix of the systems A and B are respectively
X X 0
ED 0
⇢A = j | j i h j | , ⇢B = j j j ,
j j
P
and their von Neumann entropies are S(⇢A ) = S(⇢B ) = j j log2 j.
7
n
2 1
1 X
= n |xiA |xiB .
2 2 x=0
In the above, x is taken to be the decimal number corresponding to the bi-
nary number represented by strings of n bits and A and B represent the full
collections of n qubits of the subscripts. Thus an equal amplitude superpo-
sition of d = 2n orthogonal states contains n = log2 d ebits. The entangled
state of two qudits
d 1
1 X
| d iAB = p |kiA |kiB (9)
d k=0
thus contains log2 d ebits of entanglement. We now consider how to quan-
tify the entanglement of two qubits in a general two qubit entangled pure
state. As these would result in imperfect protocols, these should have a lower
(fractional) ebit of entanglement in the units introduced here.
8
p 00 AB
+ q 11 AB
Φ+
AB
m = nS(ρ )
9
duced density matrices are given by ⇢Aj = p |0i h0|Aj + q |1i h1|Aj and ⇢Bj =
p |0i h0|Bj + q |1i h1|Bj with the von Neumann entropy S(⇢Aj ) = S(⇢Bj ) =
Q
p log2 p q log2 q). Then their combined state nj=1 |⇠iAj Bj will have, for
large n approximately np 0’s and nq 1’s on each of Alice’s and Bob’s side
(you can understand this better by constructing the reduced density matrix
on each side and then considering the law of large numbers to get the typical
sequences which are possible for large n – all other sequences have a vanishing
probability). Thus
n
Cnp
n
Y X
1
|⇠iAj Bj ⇡ pn |jiA |jiB , (10)
j=1
Cnp j=1
Exercise .4
Find out the entanglement, as quantified by the von Neumann entropy of
two systems A and B in the entangled states (a) | iAB = p12 |0iA |0iB +
p
p1 3
2
|1iA |+iB , (b) |⌘iAB = 2
|0iA |1iB + 12 |1iA |0iB .
10
Mixed State Entanglement, its Quantification
and Distillation
Sougato Bose
February 16, 2021
1
Φ+
Ai Bi
ρ Ai Bi
2
ρ Ai Bi
Φ+
Ai Bi m=
nED ( ρ Ai Bi )
3
Ψ+ = I Ai ⊗ X Bi Φ+
Ai Bi Ai Bi
Φ+
Ai Bi
1− P
Φ+
Ai Bi
4
1 Example of an Entanglement Distillation
Procedure
We will not discuss entanglement distillation in a general context, but only
illustrate its possibility through an example. Consider a particular quantum
channel in which a qubit has only a random chance of a bit flip, i.e., with
a probability P it applies a X operator on the qubit traversing the channel.
We consider Alice generating a state of two qubits | + iAj Bj and sending the
qubit Bj through the channel to Bob. Thus
+
↵
With Probability 1 P, Alice and Bob share A j Bj
,
+
↵ +
↵
With Probability P, Alice and Bob share 1Aj ⌦ XBj A j Bj
= A j Bj
.
The action of this bit-flip channel is depicted in Fig.3. We can write the
above compactly as Alice and Bob sharing the mixed state
↵⌦ + ↵⌦ +
⇢Aj Bj = P + A j Bj
+ (1 P ) + A j Bj
.
(Note that when we write mixed states in the “ket-bra” notation, we often
omit the qubit label – here Aj Bj , subscripts on the ket part, just for com-
pactness). We will perform the circuit shown in Fig.4 on two copies of such
shared states ⇢A1 B1 and ⇢A2 B2 , measure the qubits A2 and B2 as shown in
the figure, and, conditional on the coincidence outcome (when the state of
qubits A2 and B2 measured in the computational basis match), retain the
states of qubits A1 and B1 . Consider 4 cases separately:
↵ ↵
With Probability (1 P )2 : +
A 1 B1
+
A 2 B2
CN OTA1 A2 CN OTB1 B2 +
↵ +
↵
! A 1 B1 A 2 B2
5
ρ A1B1
ρ A2 B2
X X
ρ (1)
A1B1
Thus in e↵ect, the state of the systems A1 and B1 are projected, conditional
on concidence, whose total probability is P 2 + (1 P )2 , to
(1) ↵⌦ ↵⌦
⇢A1 B1 = P (1) + +
A 1 B1
+ (1 P (1) ) + +
A 1 B1
,
P2
with P (1) = P 2 +(1 P )2
. Now note that when the bit-flip probability P < 12 ,
(1)
the relative proportion of | + iA1 B1 is more in the state ⇢A1 B1 in comparison
(1)
to the initial states ⇢Aj Bj shared between Alice and Bob. Thus ⇢A1 B1 is
now a better approximation to | + iA1 B1 than ⇢A1 B1 . Many such qubit pairs
(1)
(⇠ n{P 2 + (1 P )2 }) in the state ⇢Aj Bj will thus be obtained after the above
(1)
step, which we call the first iteration. However, ⇢Aj Bj has the same form
as ⇢Aj Bj except for the probabilities P being be replaced by P (1) . Thus one
can go on repeating the same procedure iteratively on pairs available at the
(k) (k+1)
kth stage of iteration ⇢Aj Bj giving, probabilistically ⇢Aj Bj for the next stage
of iteration. For P 6= 1/2, the proportion of a single Bell state in the state
(either | + iAj Bj or | + iAj Bj ) monotonically increases with each iteration.
In this way, after quite a few iterations a state is obtained which has a very
high proprotion of a single Bell state.
6
Exercise .1
(k)
Choosing an initial state ⇢Aj Bj with P < 1/2 (make your own choice of P ,
e.g., P = 1/4), work out the number of iterations k needed to get a state with
1 P (k) 0.99. Find the yield (fraction m/n) obtained by the above distillation
procedure.
Exercise .2
Show that the above procedure can perfectly distill (convert to a Bell state)
a mixed state
↵⌦ +
⌘A j B j = P + A j Bj
+ (1 P ) |00i h00|Aj Bj
in one step, and find the probability for this one step distillation to happen.
Explain why for P > 12 , this probability is exceeded by the asymptotic (n ! 1)
fraction m/n of states arbitrarily close to a Bell states.
2 Coherent information
In general it is very difficult to find the optimal procedure for entangle-
ment distillation for an arbitrary entangled state ⇢AB . However it is known
(Devetak-Winter 2003) that if only one way classical communication between
Alice and Bob is allowed (e.g. from Alice to Bob, but not vice-versa), then
one can asymptotically distill a fraction of Bell states called the Coherent
information given by
We give here the above formula without proof. Note that for pure states
(S(⇢AB ) = 0) the fraction obtainable reduces to the von Neumann entropy
of the reduced density matrix ⇢B . For mixed states, the rough intuition is that
while nS(⇢B ) Bell states would have been achieved without the mixedness,
one has to still “learn” nS(⇢AB ) bits of information to convert the results
to a pure Bell state (to purify), which sacrifices nS(⇢AB ) Bell pairs giving a
remaining number of nIc (A > B).
7
classical communications is called a separable state (no transfer of qubits or
sharing of entanglement is allowed). This is then defined as
X
⇢Sep
AB = Pi ⌘(i)A ⌦ (i)B ,
i
where ⌘(i)A and (i)B are density matrices. Essentially we are telling that
because product (factorizable) states are the pure state versions of unentan-
gled states, probablity distributions over a number of such factorizable states
cannot be entangled either (just expand ⌘(i)A and (i)B in their eigenbasis
to check that this statement is same as the definition above). Note that these
can be correlated states. As you can see from Fig.5, Alice can throw a die
and corresponding to the outcome choose a state ⌘(i)A from an enormous
reservior of quantum states that she has, inform Bob, who correspondingly
chooses a state (i)B . Any state which can be written in the above form is
a separable state – not entangled. For example, the state
1
⇢Class,Max = (|0i h0|A ⌦ |0i h0|A + |1i h1|A ⌦ |1i h1|A ),
2
is as maximally correlated as is possible classically – Alice and Bob always
have the same state in the computational basis – but is separable, as it can
be generated by the above classical communication procedure. Any density
matrix not expressible in the form of ⇢Sep
AB above is entangled.
Exercise .3
By expressing the density matrix of an equal mixture of all four Bell states in
the computational basis, show that it is a separable state.
8
i
Sep
ρ AB = ∑ Pi η (i) A ⊗ χ (i)B
i
η (1) A χ (1)B
η (2) A η (3) A χ (2)B χ (3)B
…….
…….
Figure 5: Separable states being generated by Alice and Bob via telephone
interaction.
9
As you can see the “bra” and “ket” parts of the system B are swapped.
Similarly the partial transpose over A can be defined. This partial transpose
is not a physical operation – it is merely a mathematical operation. Now
note that X
⇢Sep
AB
TB
= Pi ⌘(i)A ⌦ (i)TB
i
⇢Sep
AB
TB
> 0 i.e., ⇢Sep
AB
TB
has positive eigenvalues.
if ⇢TAB
B
has negative eigenvalues =) ⇢AB is entangled.
Exercise .4
+ +
By using the Peres-Horodecki criterion, show that ⌘AB = P | ih |AB +
(1 P ) |00i h00|AB is an entangled state for any P 6= 0.
10
above where we mixed all Bell states (each maximally entangled) to obtain
a separable state. This is motivated by the fact that forgetting informa-
tion should only decrease entanglement. Quantities which satisfy the above
criteria are called entanglement measures.
While many such quantitites can be defined, very few are computable.
For example the distillable entanglement ED is obviously an entanglement
measure (as it is defined as the maximum fraction of Bell states extractable
by LOCC procedures, it cannot be increased under LOCC), but it cannot be
computed for arbitrary states. The one way LOCC distillable fraction given
by the Ic (A > B) is also not a measure as it could possibly enhance when
two way LOCC is allowed. A good and popular computable measure for two
qubits is called the concurrence (invented by Wootters 1998) which we will
not cover here (it is related to the amount of entanglement in the form of
Bell states needed to form a given entangled state – called the entanglement
of formation).
We will use here a measure called the negativity, which applies for arbitary
dimensional quantum systems. It is defined using the partial transpose op-
eration as follows:
X
N (⇢AB ) = | j |, where j are the eignevalues of ⇢TAB
B
.
j, j <0
Thus the modulus of the negative eigenvalues of ⇢AB are summed to get the
negativity. Armed with this simple prescription, one can now ascribe an
amount of entanglement to any mixed state ⇢AB of two systems.
Exercise .5
Compute the entanglement as quantified by the negativity for the state ⇢AB =
P | + i h + |AB + (1 P ) | + i h + |AB and find the only point at which the
entanglement is zero.
11
Quantum Algorithms 1: Illustrations of
Quantum Speedup
Sougato Bose
February 22, 2021
1
0 0 eiφ (0) 0
1 eiφ (1) 1
.
N −1 eiφ ( N−1) N −1
Hadamard transform to create Encoding the A quantum transform crea1ng Measure to find
the ini1al problem interference to solve the the answer
Large superposi1on of N=2n In terms of problem.
states of n qubits. phases.
2
Hadamard operation on each qubit, known the Hadamard transform, to be
discussed in section 6. In terms of operation time, we can regard this as
a single efficient step as all the Hadamards can be performed in parallel in
the same time-slot (when we include decoherence, to be described in a later
lecture, we will understand that we have to include the e↵ect of decoherence
in each of these n operations). The Hadamard transform leads to what we call
the starting state of all quantum algorithms with n qubits (obviously there
will be algorithms which do not start with this state, but we will mostly see
examples of those that do):
N 1
1 X
| i= p |xi , (1)
N x=0
where N = 2n and x represents all the 2n possible states of n bits, from
00..0 to 11..1. Actually we will use two equivalent representations of x when
we use the above equation, or similar equations, in various contexts. x will
sometimes mean the bit string x = x1 x2 x3 ...xn (where each xj 2 {0, 1}
is a bit), and will some times be used to mean the decimal number x =
x1 2n 1 + x2 2n 2 + ... + xn 1 2 + xn – it will be clear from the context which
way we are using x in a certain step of an equation.
Another crucial fact to note is that, although we are using the standard
multi-slit interferometry for the depiction of a generic quantum algorithm,
there are a number of diferences. The interference is not in a whole continuous
variable spatial degree of freedom as in standard interferometry learnt in
physics, but in a discretized version of it – only superpositions of the states
|xi , x = 0, 1, .., 2n 1 occur at any time during the algorithm – it is as if space
was discretized and finite with only the above values of x allowed. Moreover,
the transform need not be the Fourier transform which occurs in standard
interferometry, it could be any other appropriately chosen transform which is
useful to solve the problem (although we know that the Fourier transform is
used in some of the most powerful quantum algorithms). What do we mean
by a transform here?P It is a unitary evolution which evolves any state |xi to
a superposition k cx,k |ki with non-zero complex amplitudes cx,k for several
k 6= x} (in other words, the evolution is non-diagonal in the |xi basis).
3
Classical
x BB for f(x)
f(x)
(a)
x A Quantum x A
BQBB for
b B
f(x) b ⊕ f (x) B
(b)
x Quantum (−1) f ( x ) x
A
BQBB for
A
Quantum
1
2
(0 B
−1 B ) f(x)
1
2
(0 B
−1 B ) ≡ PQBB for
f(x)
(c)
where the modular addition guarantees the orthogonality of outputs for or-
thogonal input states of the ancilla B.
However, as described in the last section, for several quantum algorithms,
4
the function f (x) is encoded in the problem in terms of phases (x) imparted
to the states |xi. Thus we will need a way to convert the information given
as a bit string (more precisely as a string of qubits in computational basis
states) to an information given as a phase acquired by the |xi states. Let
us temporarily restrict to f (x) 2 {0, 1} as we will only need this case for
the algorithms of this lecture. Then the conversion to a phase is done most
conveniently by assuming the ancilla to be a single qubit, and feeding it to
the BQBB in the state | iB = p12 (|0iB |1iB ). The evolution of the BQBB
with this input is thus (Fig.2(c))
1 BQBB 1
|xiA p (|0iB |1iB ) ! p {|xiA |0 f (x)iB |xiA |1 f (x)iB }
2 2
1 ¯ ↵ }(where f (x)
= p {|xiA |f (x)iB |xiA f (x) ¯ stands for the complement of f (x);
B
2
¯ = 1, f (x) = 1 =) f (x)
f(x)=0 =) f (x) ¯ = 0)
1
= |xiA p (|0iB |1iB )for f (x) = 0;
2
1 1
= |xiA p (|1iB |0iB ) = |xiA p (|0iB |1iB )for f (x) = 1.
2 2
The above two steps can be summarized as
BQBB
|xiA | iB ! ( 1)f (x) |xiA | iB .
5
f (0), and the other by inputting x = 1 and getting f (1). Then we can com-
pare f (0) with f (1) to conclude whether the function is constant or balanced.
6
Classical Algorithm: The natural strategy will be to take each value of
x one by one and use the oracle to evaluate the corresponding f (x). As soon
as we find a f (x) which has a di↵erent value from all the previous evalua-
tions, we infer that the function is not constant (i.e.it is balanced). It is quite
intuitive that in the worst case you will require N2 + 1 evaluations of f (x) to
ascertain whether it is constant or balanced. This is because all the first N2
evaluations may unluckily coincide even if the function is balanced, but the
N
2
th evaluation has to show a di↵erence. This is an exponential number of
evaluations as N = 2n .
It is easy to verify (satisfy yourself by thinking about this for a moment) that
from the condition on the function (either ↵constant or balanced) it follows
that h | ? i = 0, and thereby | i and ? can be fully distinguished in a
quantum measurement. Thus a single usage of the PQBB has sufficed to
distinguish between a constant and a balanced function irrespective of the
case – works equally well in all possibilities of the function – no best case
or worst case. Thus over the classical algorithm’s worst case, the quantum
algorithm has a N2 + 1 : 1 speedup or an exponential speedup.
Exercise .1
Show that the starting state of all quantum algorithms can be generated as
follows by acting by a Hadamard on n qubits, each in a state |0i, i.e., show that
N 1
⌦n 1 X
H |00..0i = | i = p |xi .
N x=0
6 Hadamard Transform
The Hadamard transform is one of the unitary operations which is used in
some quantum algorithms to implement the giant multiparticle interferom-
etry over the 2n states of n qubits. It is also the unitary operation used to
generate the large equal superposition over all the 2n states in the first place.
7
Let us recall that
1
H|0i = p (|0i + |1i) (4a)
2
1
H|1i = p (|0i |1i) (4b)
2
The above two equations can be written in a compact form as
1 X
H|xi i = p ( 1)xi yi |yi i (5)
2 y 2{0,1}
i
1 X X X
= n ... ( 1)x1 y1 +x2 y2 +...+xn yn |y1 y2 ...yn i
2 y 2{0,1} y 2{0,1} y 2{0,1}
2
1 2 n
N 1
1 X
= p ( 1)x1 y1 +x2 y2 +...+xn yn |y1 y2 ...yn i
N y=0
N 1
1 X
= p ( 1)x1 y1 x2 y2 ... xn yn
|y1 y2 ...yn i (6b)
N y=0
N 1
1 X
= p ( 1)x.y |yi, (6c)
N y=0
8
course, we will assume, as discussed in section 3, that the Black-Box operates
as a phase query Black-Box (PQBB) in the quantum algorithm. The task at
hand is described below:
Task: Find a.
Quantum Algorithm: PWe1 start with the usual starting state of quan-
tum algorithms | i = p1N N
x=0 |xi. This, as shown before, is easily gener-
⌦n
ated by H on the qubit state |00...0i.
N 1 N 1
1 X PQBB evaluating a.x 1 X
p |xi ! p ( 1)a.x |xi
N x=0 N x=0
N 1N 1
H ⌦n 1 XX
! ( 1)a.x ( 1)y.x |yi (7a)
N x=0 y=0
= |ai. (7b)
How does the astonishing equivalence between Eq.(7a) and Eq.(7b) follow?
To understand that, consider the amplitude of the state |y = ai in Eq.(7a):
N 1
1 X
( 1)a.x ( 1)a.x = 1. (8a)
N x=0
As the entire quantum algorithm uses unitary operations only, the normal-
ized input state | i must have evolved to a normalized output state at the
end of the algorithm. As the amplitude of the state |y = ai is unity, and it
is itself a normalized state, thus the amplitude of all states |y 6= ai must be
9
zero in the output state. Thus the output state must be |ai (i.e. Eq.(7a)
is equivalent to Eq.(7b)). Eq.(7b)) implies that we can simply readout the
value of a from the output state. In all this procedure, there has been only
one application of the Black-Box, as a phase query black box evaluating a.x.
Thus,
Task: Find a.
10
of a collision between any two evaluations is exponentially small (because x
and x a can be any two arbitrary values among an exponentially large set
of N = 2n possible values of x). Thus classically, we have to have an expo-
nentially large number of queries in the average case to determine the string a.
Eq.(8ba) implies that if we measure the first register now and find a bit string
y = y (1) (where we have used superscript and brackets to indicate that y (1)
is a bit string), then this bit string y (1) satisfies the linear equation
Now one repeats the whole algorithm from the start O(n) times and obtains,
by measuring the first register at the end, another set of distinct bit strings
y (2) , y (3) , ...., y (n) so that
In the case that one of the y (j) s coincide with the outcome of a previous
measurement (whose chance is exponentially small anyway) the algorithm
must be repeated again to obtain another distinct y (j) . In this way, by
O(n) repeatitions, one can obtain n linearly independent linear equations,
Eqs.(8ca)-(8df), from which a can be obtained. Thus,
11
Speedup: The quantum algorithm, which uses O(n) runs of the algo-
rithm (hence O(n) uses of the oracle) achieves an exponential speedup over
n
its classical counterpart which requires O(2 2 ) queries. This is an exponential
speedup, and it is not just a speedup over the worst case but a speedup in
the average case.
References
[1] E. Bernstein and U. Vazirani, Quantum complexity theory, SIAM Jour-
nal on Computing, 26(5):1411-1473, 1997.
12
Quantum Search & Introducing Quantum
Fourier Transform
Sougato Bose
March 2, 2021
1
states, such as a solution to a Boolean satisfiability problem, for instance
using quantum methods. These can then be amplified by the methodology of
Grover’s search algorithm. Thus it has an important usage as a subroutine
of various quantum algorithms.
Exercise .1
Verify that the action of the above PQBB is given by the operator
i.e.,
Ôm |xi = |xi if x 6= m,
Ôm |xi = |xi if x = m.
We will discuss the implementation of the database and a BQBB version
of the above quantum oracle in the next section (essentially a quantum circuit
which notifies the match of an input bitstring x with a marked bit string m),
as well as a version where f (x) is the solution to a decision problem.
Secondly, the algorithm requires another unitary operator
Ô = 2 | i h | 1,
2
PN 1
where | i = p1 |xi. The unitarity of Ô is proven in exactly the same
N x=0
manner as that of Ôm above. Grover’s search algorithm is implemented
iteratively by repeatedly applying the operator Ô Ôm on to the quantum
register (meaning will become clear below).
Exercise .2
Show that
H ⌦n Ô0 H ⌦n = Ô .
The above exercise shows that Ô can, in turn, be implemented by a
PQBB with marked element x = 0, i.e., Ô0 .
Quantum Algorithm:
The quantum algorithm is simply
p the application of Ô Ôm on the state
1
PN 1
| i = pN x=0 |xi about O( N ) times. In other words, the claim is
Exercise .3
Making your own choice of a marked element (any of 0, 1, 2 or 3) among four
possibilities, show that a single application of the Ô Ôm operator implements
the Grover search perfectly for a 2 qubit register (i.e., evolves the state exactly
to the marked element).
with
N
X1
?
↵1
m =p |xi.
N 1 x=0,x6=m
↵
Note that hm? |mi = 0. Thus {|mi , m? } can be used as a complete basis in
a 2-dimensional Hilbert p and the state | i makes an
↵ space as shown in Fig.1
angle ✓ with the m? such that sin ✓ = 1/ N . In this picture, the actions
of the two operators are
↵
Ôm reflects states about the m? axis,
3
m k
(
ψ (k ) = Ôψ Ôm ) ψ
2
(
ψ (2) = Ôψ Ôm ) ψ
2θ
2θ ψ
θ m⊥
θ
2θ Ôm ψ
Ôm ψ (1)
Figure 1: Figure showing the angle rotations due to each application of the
grover operator Ô Ôm . Note that the angles and the lengths of the vectors
as shown in the figure are not to scale.
(as the amplitude of the state |mi becomes negative while keeping the same
magnitude)
Ô reflects states about the | i axis,
(as the amplitudes of all states other than | i becomes negative while keep-
ing the same magnitudes).
Using the above actions, the evolution of the angles during the interations
of Ô Ôm take place as follows (read these lines while comparing with Fig.1):
↵
Ôm | i makes an angle ✓ with m? ,
↵ ↵
(1)
= Ô Ôm | i makes an angle 3✓ with m? ,
↵ ↵
Ôm (1) makes an angle 3✓ with m? ,
↵ ↵
(2)
= (Ô Ôm )2 | i makes an angle 5✓ with m? ,
.
.
.
↵ ↵
(k)
= (Ô Ôm )2 | i makes an angle (2k + 1)✓ with m? .
4
Thus each iteraction Ô Ôm rotates the state by an extra angle 2✓ anticlock-
wise starting from the initial state | i. When the rotation angle reaches ⇡/2,
the rotated state is approximately |mi and our algorithm is finished (marked
state has been reached). The number of iterations k required for this is
⇡
(2k + 1)✓ ⇡ .
2
If N is large, ✓ ! 0, sin ✓ ⇡ ✓ = p1 . Thus we have
N
1 ⇡
(2k + 1) p ⇡
N 2
⇡p
=) k ⇠ N.
4
p
Thus, in k ⇠ O( N ), iterations of Ô Ôm (usages of the oracle), the ampli-
tude of the marked state, which was initially p1N has been amplified to near
unity (this is thus also often called amplitude amplification), and we would
find the marked state with a very high probability if we measure.
p
Speedup: Thus Grover’s search algorithm o↵ers a O( N ) speedup over
classical algorithms in finding a marked element from an unsorted database.
5
a1 a1
a2 a2
x1 x1
x2 x2 If
If
f(x)=1
m1 m1 meas
a1 a2
m2 m2
0 X X X X 0
0 X X X X
0
0 X f (x)
Figure 2: Figure showing the construction of Ôm (qubits 3rd to 9th from the
top). Two qubits are added to the top to make the first 4 qubits a quantum
database e.g, name, phone number combinations.
How does it fit in with the original motivation of searching over an un-
sorted database? Let us thus extend our simple example above to an un-
sorted database encoded in quantum states. Each entry to the database
is |a1 a2 x1 x2 i. Consider, as above, only 4 entries to the database, but now
labelled by two bit strings a = a1 a2 (these are essentially equivalent to the
names of the people in the database), and two bit phone numbers represented
by strings x = x1 x2 . We know a specific phone number m = m1 m2 , but do
not know the “name” of the person a1 a2 to whom the number m belongs. Our
task is to find a1 a2 using a quantum algorithm – so we construct a quantum
oracle as shown in Fig.2. Note that the |a1 a2 i part of the input state does
not seem to have been used for anything in the circuit. It just maintains the
correspondence of names with phone numbers. When a match of the phone
number is found i.e., f (x) = 1 is obtained, one merely measures them to find
out the name of the person to whom the phone number m belongs. To clarify
a bit more, in the usual starting state of all quantum computation | i, each
state |xi would actually stand for |a(x), xi where a(x) is the bit string that
gives the name of p the person who has the telephone number x. When |mi is
amplified after O( N ) search iterations, one can simply read o↵ |a(m)i.
The above usage, the literal search of an unsorted database is, of course,
very difficult precisely because of the need to prepare the quantum database
and the initial state | i including the information of the names a(x) to which
each x belongs. Although the circuit for Ôm is easy, as shown above, and so
6
will the circuit for Ô0 , but the circuit for Ô may not be easier in this case.
Thus the quantum search algorithm, in the sense of amplitude amplifica-
tion is more useful for cases where the oracle gives |f (x) = 1i for the affirma-
tive instance of a decision problem, otherwise |f (x) = 0i. Here the entries are
solely the bit strings |xi with no extra labels such as name entries, so that the
algorithm described in the previous section can be implemented just as de-
scribed. In these circumstances, the amplitude of the bit strings |xi for affir-
mative case/s can be amplified by the quantum algorithm and then measured
to find those x with give f (x) = 1. For a neat example of this oracle construc-
tion for 3-SAT, please see question, answer and circuit (see all 3 circuits as
they progressively build) at: https://cstheory.stackexchange.com/questions/38538/oracle-
construction-for-grovers-algorithm
The 3-SAT problem itself has a best known exponential O(K n ) classical
complexity (K > 1 is some base, which decreases with better the algorithm
you find), which can be reduced to O(K n/2 ) by the application of the square
root speedup given by amplitude amplification. In short, any classical oracle
computing a function f (x) efficiently for any given entry x is, in the end made
from a circuit composed of AND, OR, NOT. These can all be replicated as
classical reversible (and hence quantum) circuits build from Fredkin and/or
To↵oli gates. Then, if it is a decision problem with f (x) 2 {0, 1}, we can
always convert it to a PQBB to give a ( 1)x phase to those x entries which
give f (x) = 1. Then those values of x can be thought as marked elements
(marked by the decision problem you are solving), which can all be amplified
by Grover’s amplitude amplification algorithm.
7
0
2
1
3
4
Figure 3: A fully connected graph of sites on which a single particle can hop,
along with a defect site |2i of di↵erent energy.
Hw = Hd + Hp = | i h | + |mi hm| .
8
p
time t = ⇡2 N . Essentially the problem can again be solved by converting it
to an e↵ective problem ↵in a 2-dimensional Hilbert space with states expressed
in the basis {|mi , m? }. Thus without knowing which vertex has di↵erent
p
energy apriori, quantum mechanics can find that vertex in O( N ) time.
Exercise .4
q q
↵
N 1 ? 1
Using | i = N
m + N
|mi and writing the Hamiltonian in the
↵
{|mi , m? } basis, show that the minimum gap of Ha (t) is p2N .
Duepto the result of the above exercise, we can conclude that in time
t ⇠ O( N ) one can reach the marked state |mi via the quantum adiabatic
algorithm.
The above implementations are often called a spatial search for the ob-
vious reasons (e.g. see Fig.3). We will next proceed to a di↵erent class of
quantum algorithm, which gives an exponential speedup over known classical
algorithms. For that we require to use the quantum fourier transform, which
we introduce in this lecture, with its algorithmic applications discussed in
the next lecture.
In the above, |ki are also bit-string states as defined for j above, jk means
straightforward multiplication of the decimal values of j and k (i.e. j1 2n 1 +
j2 2n 2 +p
... + jn and k1 2n 1 + k2 2n 2 + ... + kn respectively), i is the complex
number 1, and, as usual N = 2n where n is the number of qubits.
5.1 Exercise:
PN 1 PN 1 i 2⇡ jk
Show that F = p1 |kihj|. Hence prove that F is a unitary
k=0 e
N
N j=0
operation.
9
5.2 Exercise:
Show that F|110i is a product state of 3 qubits.
10
Quantum Fourier Transform – Circuit &
Applications
Sougato Bose
March 9, 2021
In this lecture we will first demonstrate that the Quantum Fourier Trans-
form is an easy unitary operation to implement. We will then describe its
classic applications – Shor’s algorithm for factorization of a large number to
its primes, as well as quantum phase estimation or eigenvalue estimation (the
latter being a problem in the 3rd problem sheet, we will skip over some of
the mathematics).
1
j1 H RH22 R2H3 R2H4 4
j2 H RH22 RH
23 3
H RH22 2
j3
j4 H 1
2
in ascending order from top to bottom, while the output to the circuit is
the factorizable state of Eq.(1) with qubits labelled as 1, 2, ...n in descending
order from top to bottom (reverse of the order of the qubit labels in the
P j1 j2 jn
input). Essentially, the circuit evolves |j1 i to 1kn =0 ei2⇡( 2 + 22 +..+ 2n )kn |kn i,
P j2 j3 jn
|j2 i to 1kn 1 =0 ei2⇡( 2 + 22 +..+ 2n 1 )kn 1 |kn 1 i, and so forth. To understand the
operation of the circuit consider, in Fig.1, the conversion of |j1 i through the
circuit (where the output qubit states are labelled as k4 = {0, 1}),
1
X 1
X
H j1
j 1 k4
|j1 i ! ( 1) |k4 i = ei2⇡ 2 k4 |k4 i
k4 =0 k4 =0
1
X
CR22 j1 j2
! ei2⇡( 2 + 22 )kn |k4 i
k4 =0
1
X
CR23 j1 j2 j3
! ei2⇡( 2 + 22 + 23 )k4 |k4 i
k4 =0
1
X
CR24 j1 j2 j3 j4
! ei2⇡( 2 + 22 + 23 + 24 )k4 |k4 i .
k4 =0
Thus for each qubit state of the output, there is one block (dotted box
in Fig.1), with the number of operations in each block depending on how
many qubits are below that qubit in the circuit. Considering the general n
version of the circuit in Fig.1, we will thus have l operations (including the
H) for the evolution of |jn l+1 i to the output qubit l. Thus the total number
of operations needed for a n qubit Quantum Fourier Transform (QFT) is
n(n+1)
2
⇠ O(n2 ), which is a polynomial scaling in n implying that it is an
easy unitary operation to implement.
3
mod N =1,
r r
=) (a 2 + 1)(a 2 1) mod N = 0.
In other words, there is a potential for finding prime factors of N of the form
r
a 2 ± 1. It is very easy to check whether something is a factor by using the
gcd algorithm.
Exercise .1
Choosing N = 15 to factorize, choose a = 2 (coprime with N ) and compute
(and plot, if you think that helps you to picturize) f (x) = ax mod N for x =
0, 1, ..., 15 to show that f (x) has a period of r = 4 and is monotonic (continually
r
increases) within one period. Hence show that a 2 ± 1 finds the prime factors.
So which part of the above is hard classically? It is not computing the
function f (x) = ax mod N as powers can be very easily done if you have
multiplier circuits, which in turn, require adder circuits (you compute a.a =
a2 , then a2 .a = a3 etc). It is, in fact, finding the period of the function.
For large domain of x, when N = 2n is exponentially large in n, typically r
can also quite (exponentially) large, so if we find a f (x0 ), finding its partner
f (x0 + r) can be very difficult.
Exercise .2
Verify that for computational basis inputs, the circuit of Fig.2 accomplishes
the SUM and CARRY operations of binary addition in its second and third out-
puts.
Note we required 2 fundamental gates for the half adder (if the To↵oli
is designed in terms of fundamental 2 qubit gates, then slightly more) for
adding two bits. Adders for 2 n-bit string states can also be constructed as
similar circuits and require O(n) gates.
Exercise .3
Check for yourself that usual long multiplication/column multiplication rule
learnt long ago in your primary school days also works for binary numbers, e.g.,
4
a a
b
X SUM = a ⊕ b
CARRY = ab
0 X
Figure 2: Circuit for Quantum Half Adder.
just square a number 1101 (13 in decimal numbers) to get the binary number
for 169.
Using the usual long multiplication rule, multiplying any n-bit number
a = a1 a2 ...an with another, say, b = b1 b2 ...bn should thus take n2 bit by bit
(bn with an , then bn with an 1 and so on till bn with a1 , and then repeat
starting with bn 1 etc.) multiplications (To↵oli gates with the target going
to state aj bk ) and then another O(n2 ) operations for adding the columns.
Totally, thus multiplying two n-bit numbers requires O(n2 ) multiplications.
There are alternative quantum circuits for multiplying, but still requiring
O(n2 ) operations. We are being a bit loose here in that we are skipping the
fact that it is also possible to accomplish the mod N multiplications with
O(n2 ) operations – such a multiplier is called a modular multiplier (making
multipliers modular additionally require subtractors, which are adders ran
backwards, and a check of overflows).
Then we use the formula that for a bit string x = x1 x2 ...xn ,
n n 1
ax mod N = (a2 modN )x1 (a2 modN )x2 ...(amodN )xn .
Each of the factors in the above decomposition are accomplished by con-
ditional mod N multipliers (conditional on state xj ) as shown in the circuit
is given in the Fig.2 of https://arxiv.org/pdf/1207.0511.pdf (that figure de-
picts the full Shor’s algorithm; the modular exponentiation is accomplished
by the part with controlled modular multiplier gates between the H ⌦n and
inverse QFT). Firstly note that in going from modular multipliers to con-
trolled modular multipliers the circuit complexity is still of the same order
(as you have seen in the quantum circuits lecture, it is simple to go from uni-
taries to controlled unitaries). So each of the controlled modular multipliers
require O(n2 ) operations. However, there are n such controlled modular mul-
tipliers, one for each factor of the above expression for ax mod N , so totally
O(n3 ) operations to achieve a black-box circuit (say, a bit query black box)
for computing f (x) = ax mod N .
5
4 Shor’s Algorithm 1995
As we have explained above, Shor’s algorithm operates by first converting the
number theoretic porblem of factorizing a large number N into two primes
to choosing a number a = 2, ..., N 1 coprime with N and finding the period
(or order) r of a function f (x) = ax mod N . This function is monotonic
inside each period as you would have seen in the exercise example above.
It classically a hard task as after one has found f (x0 ), he/she has to find a
f (x0 + r) among an exponentially large number of possibilities.
From here on, we will thus think of the problem as one of finding the
period r of a function defined over a large domain. We may draw intuition
from the fact that quantum mechanics of a single particle acomplishes Fourier
transforms quite naturally. For a freely propagating particle, position kets
becomes their quantum fourier transforms (momentum kets) in infinite time.
Thus imagine the output of a di↵raction grating with slits separated by an
unknown distance r. This is a superposition of equally position separated
slits |x0 i + |x0 + ri + |x0 + 2ri + ... +↵|x0 + (N 1)ri, which is, in turn, a
superposition of momenta states m Nr (m = 0, 1, .., N 1). Unfortunately,
if you think of the position basis as the analog of the computational basis
(we generally always measure positions!), you cannot measure any of the
above momenta states to find m Nr directly. At a far distance from the slits,
after the quantum fourier transform, these momenta states are converted
to position states. You can then simply measure them to find m Nr and,
knowing N , deduce r from it. This is exactly how the Shor’s algorithm
works, albeit, the position basis, which is not exponentially large, is replaced
by the exponentially large computational basis, and the QFT on a perodic
superposition as above is accomplished by the circuit of Fig.1.
Let us now discuss Shor’s algorithm for period/order finding below.
Task: Find r.
6
P 1
the qubit state |00...0i) | i = p1N N x=0 |xi, while the second register (also
n-bit long) is in the state |0i = |00...0i. Then the algorithm proceeds as
follows:
N 1 N 1
1 X BQBB evaluating f (x) 1 X
p |xi|0i !p |xi|f (x)i
N x=0 N x=0
Measure 2nd register to find f (x = x0 ) 1 N
!p (|x0 i+|x0 +ri+...+|x0 +( 1)ri)|f (x0 )i.
State re-normalized after Meas. N/r r
Then,
1 N QFT circuit applying F to first register
p (|x0 i + |x0 + ri + ... + |x0 + ( 1)ri) !
N/r r
N 1
1 X i 2⇡ x0 k
1 2⇡ 2⇡ N
= | final i = p p {e N + ei N {x0 +r}k + ... + ei N {x0 +( r 1)r}k
}|ki
N/r N k=0
p N 1
N
1
r X i 2⇡ x0 k Xr
2⇡
= e N { ei N lrk } |ki
N k=0 l=0
r 1
1 X N
=p m .
r m=0 r
Speedup Only a few repeat runs of the quantum algorithm, each requir-
ing one QFT (O(n2 ) gates) and one query to the BQBB (O(n3 ) gates), it has
been possible to find r. In the classical algorithm about O(r) = O(2p ) appli-
cations of the function evaluating black-box will be required. Thus there is
an exponential speedup.
7
we can use QFT to estimate to any desired precision. A phase is define
modulo 2⇡. Suppose we want to estimate the phase to a precision 2⇡/N (in
units of 2⇡/N ), where N = 2n is very large. Say, in these units, = 2⇡ N
j,
where di↵erent values of are given for di↵erent j = 0, .., N 1. Then, we
P first
run a circuit on a register of n qubits prepared in the state | i = p1N k |ki
l
with several controlled unitaries CU 2 (with l = 0, 1, ..n) going from the
qubits of this register to a second target register. Let us call the Unitary
enacted by this circuit as V . After providing enough hints in the problem,
designing the circuit for V will be given as a problem in problem sheet 3 such
that it performs
V
| i | i ! | ( )i | i
where
N 1
1 X i k
| ( )i = p e |ki .
N k=0
Once V is accomplished, we use the inverse quantum Fourier transform op-
erator F † on the first register. From the definition of F given in the last
lecture, and substituting = 2⇡ N
j we get
F † | ( )i = |ji .
X F † on first register
X
cm | ( m )i | mi ! cm |jm i | mi .
m m
8
Decoherence – The main enemy of Quantum
Computation
Sougato Bose
March 10, 2021
HQ,j = gj ZQ ⌦ Zj ,
1
E6
g6 E1
g5
E5 g1
Q g2
g3 E2
g4
E4 E3
Figure 1: A central qubit in a spin bath.
above state of Q is
1 1 e i ei
⇢Q (0) = |0i h0|Q + |1i h1|Q + |0i h1|Q + |0i h1|Q ,
2 2 2 2
from which the phase is fully retrievable. We can say that can be read o↵
from the o↵-diagonal terms in the density matrix. Let us assume the bath
spins to be initially in the state
n
Y
|⇠(0)iE1 E2 ...En = |+iEj .
j=1
2
Thus, final state of the joint system
iHt
| (t)iQE1 E2 ...En = e | (0)iQE1 E2 ...En
Yn
1 1 igj t
= p |0iQ p (e |0iEj + eigj t |1iEj )
2 j=1
2
Yn
ei 1
+ p |1iQ p (eigj t |0iEj + e igj t
|1iEj )
2 j=1
2
1
= p (|0iQ |⇠0 (t)iE1 E2 ...En + ei |1iQ |⇠1 (t)iE1 E2 ...En ),
2
where we have defined
Yn
1 igj t
|⇠0 (t)iE1 E2 ...En = p (e |0iEj + eigj t |1iEj ),
j=1
2
Yn
1
|⇠1 (t)iE1 E2 ...En = p (eigj t |0iEj + e igj t
|1iEj ).
j=1
2
Note that as |⇠0 (t)iE1 E2 ...En 6= |⇠1 (t)iE1 E2 ...En , the state | (t)iQE1 E2 ...En is not
factorizable between the qubit and its environment. Thus, due to the Ising
interaction, the qubit has become entangled with its environment. This en-
tanglement leads to the phenomenon of decoherence. From earlier lectures
on reduced density matrices, you would thus immediately note (do this ex-
ercise by yourself and verify) that the reduced density matrix of the qubit at
a time t is
1 1 e i ei
⇢Q (t) = |0i h0|Q + |1i h1|Q + |0i h1|Q h⇠1 (t)|⇠0 (t)i+ |0i h1|Q h⇠0 (t)|⇠1 (t)i.
2 2 2 2
Exercise .1
Qn
2igj t 2igj t Qn
Show that h⇠0 (t)|⇠1 (t)i = j=1 e +e 2
= j=1 cos 2gj t.
Thus is matrix form, the density matrix is given by
" Qn #
ei
1 1 2 j=1 cos 2g j t
⇢Q (t) = i Qn .
2 e2 j=1 cos 2g j t 1
Now let us examine the behaviour of the o↵ diagonal terms of the density
matrix. If there was a single environmental spin (n = 1), the o↵-diagonal
terms of the density matrix would initially decrease in magnitude as a cosine
with time, and become zero at t = 4g⇡1 t . Thus at this particular instant of
time, the density matrix has lost all its information about as ⇢Q (t = 4g⇡1 t )
has no in its expression! However, this single environmental spin case is
not a true decoherence, as the o↵ diagonal terms containing start reviving
as soon as t increases, reaching the original magnitude again at t = 2g⇡1 t ,
albeit becoming negative. However, for very large n, the magnitude of the
3
o↵ diagonal terms of the density matrix are less that 12 for any general time
t 6= 0 as long as the interaction strengths gj are incommensurate (they cannot
all revive at the same time!) and quite uniformly and densely distributed.
Moreover, the o↵ diagonal terms, being a product of many cosines of incom-
⇡
mensurate frequencies, are typically very small after t 4gj t
for the fastest
frequency gj . As t ! 1, it is unlikely that at any instant of time all of
the cos 2gj t terms will become large enough to give a significant non-zero o↵-
diagonal term – for several frequency/frequencies they will remain very small
(including exactly zero for some), and moreover, it is a product of several
such terms. So we can safely say,
" #
1 ei 1
1 t!1 1 0
⇢Q (0) = e i
2 ! ⇢Q (t) = .
2 2
1 2 0 1
The above phenomenon of the o↵-diagonal terms going to zero is called de-
coherence. Note that has completely disappeared from the state – so any
quantum algorithm with a problem encoded in for example, or the di↵er-
ence between the states |+i ( = 0) and | i ( = ⇡) in QKD for example, or
the di↵erence between | + i and | i in quantum dense coding or teleporta-
tion for example, will not work any more. Decoherence is thus the principal
enemy of quantum information processing.
Exercise .2
Repeat the above
Qn calculations for environmental spins in the initial states
⇢(0)E1 E2 ...E n = j=1 (|0i h0|Ej + |1i h1|Ej ), which can be thought of as an
infinite temperature state of the environment. Give a clever logic as to why the
calculations give exactly the same result as in the case we have considered above.
4
neglected interaction terms) when the environment itself is well isolated from
its surroundings, such as the nuclear spin bath of an electronic spin in a
nanocrystal – a realistic initial state of the environment is that of the exer-
cise problem rather than the case we have worked out above. Here the phase
information is contained in the qubit-environment system under considera-
tion, and not leaked out to a larger environment outside – the considered
environment is hardly interacting with its wider environment outside over
the time scales considered. Such an environment is an example of a non-
Markovian environment. In our model, for example, at short times, the o↵
diagonal terms decay as
n
Y n
Y Pn
t!0 2 2 2
cos 2gj t ! (1 2gj2 t2 ) ⇠ e j=1 gj t .
j=1 j=1
5
Exercise .3
Show that ⇢Q (2t) = ⇢Q (0) in the toy qubit-environment model studied by us
if at time t an operation XQ is applied to Q.
The above procedure is called dynamical decoupling and is widely used in
preventing dephasing (the type of decoherence considered here, where phase
information is lost) in various qubits. Repeated pulses are given at regular
intervals to delay the dephasing. In contrast to our toy model, realistically,
the dephasing can only be delayed and not perfectly reversed as informa-
tion still gradually leaks from our environment of spins directly interacting
with the qubit to their own wider environments and, additionally the actions
of noncommuting terms in the environmental spins, including their interac-
tions with each other, make even a partial cancellation more difficult as time
progresses.
Exercise .4
Show that for Markovian decoherence as described above,
t t
1+e 1 e
⇢Q (t) = ⇢Q (0) + ZQ ⇢Q (0)ZQ
2 2
.
The above result has implications on how to treat decoherence on a qubit
in a discrete way. We say that at any time (or in any process that takes
some time, such as a quantum gate or the propagation of a qubit through
a quantum channel or a qubit simply sitting in a memory) the process of
dephasing can be described in terms of probabilities of a “phase flip” (the
application of a ZQ operator). System evolves as
⇢Q ! (1 P )⇢Q + P ZQ ⇢Q ZQ .
The above is said to be the action of a phase-flip quantum channel on a
qubit. While we had a model with ZQ ZEj interactions, it is easy to imagine
a model with other interactions e.g., XQ XEj , YQ YEj etc. Thus we also have
the possibilities of a “bit-flip” channel
⇢Q ! (1 P )⇢Q + P XQ ⇢Q XQ ,
and both bit and phase flip combined channel
⇢Q ! (1 P )⇢Q + P YQ ⇢Q YQ .
When a polarized photon encoding a qubit passes through a fiber, all the
above 3 processes could occur with equal probabilities. We then say that we
have a depolarizing channel given by
P
⇢Q ! (1 P )⇢Q + (XQ ⇢Q XQ + YQ ⇢Q YQ + ZQ ⇢Q ZQ ).
3
6
How to counter such forms of decoherence when it is a↵ecting quantum com-
munication, was discussed when we studied entanglement distillation. How
to counter such decoherence in a general setting in any quantum information
process, particularly quantum computation, is called quantum error correc-
tion, which will be discussed in the next lecture.
7
Quantum Error Correction
Sougato Bose
March 16, 2021
1
To clarify, we just choose an orthogonal basis of the environment |⇠k i and
write the joint state of Q and E in terms of them by taking them out as com-
mon factors and grouping the states of Q that correspond to them. These
states of Q corresponding to each |⇠k i are then normalized (these represent
states | k i, p
which are, in general, non-orthogonal) so as to yield appropriate
amplitudes Pk ei✓k . In the lower equation, we have then introduced unitary
matrices U (k) which map the normalized state | iQ to the various normal-
ized states | k i (i.e., U (k) | iQ = | k i). Thinking about the system only (as
the environment is inaccessible) the above evolution can be written in terms
of the reduced density matrix of the qubit as follows
X
⇢Q = | i h | Q ! Pk U (k)Q ⇢Q U (k)†Q .
k
2
Of course, you require convincing that the above type of syndrome measure-
ment, collapsing a state to something such as the outcome Xj ↵ |0L i12...n +
Xj |1L i12...n , in which the information inherent in the quantum state in
terms of ↵ and is still preserved, is possible. For that it is best to see
an explicit example of a quantum error correcting code – an explicit exam-
ple of |0L i12...n and |1L i12...n with the accompanying syndrome measurement
strategies.
Exercise .1
Show that the circuit of Fig.1 converts |0i1 to |0L i and |1i1 to |1L i, from
which is follows that it will convert ↵ |0i1 + |1i1 to ↵ |0L i + |1L i.
In essence, the code is constructed as above so that a single bit flip error
can be detected via a majority voting among changes to |000i and |111i,
while a single phase flip error can be detected via a majority voting among
the phases ±1 between |000i and |111i in the 3 brackets. Thus the syndrome
measuring operators are
Exercise .2
Verify that |0L i and |1L i, and hence ↵ |0L i + |1L i are eigenstates of M (j)
with eigenvalues +1.
Let us check directly the action assuming a bit-flip error on the 5th phys-
ical qubit (X5 )
3
α 0 +β 1 H
0 X
0 X
0 X H
0 X α 0 L + β 1L
0
X
0 X H
0 X
0 X
4
X5 (↵ |0L i+ |1L i) = ↵ p13 (|000i+|111i) p13 (|010i+|101i) p13 (|000i+|111i),
1 1 1
+ p (|000i |111i) p (|010i |101i) p (|000i |111i).
3 3 3
Using the right hand side of the above equation, we find
Now consider a phase flip error on the 7th qubit (Z7 ). Then, it is easy to
verify that
M (1) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (2) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (3) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (4) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (5) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (6) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (7) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i),
M (8) Z7 (↵ |0L i + |1L i) = Z7 (↵ |0L i + |1L i).
Thus a 1 eigenvalue of M (8) only (+1 for all the other syndrome operators)
indicates that the error was either Z7 or Z8 or Z9 . Either of these operators,
say, Z8 can be then applied to correct the state
5
Exercise .3
Find the combination of syndrome operator M (j) eigenvalues to detect a X3
and a X6 Z6 error, and which operators can be applied to correct these errors.
We have shown above that the eigenvalues of the syndrome operators can
help you identify the specific error. But how does one actually measure these
eigenvalues without destroying the state? For example, measuring (say, by
a Stern-Gerlach for physical qubits which are spins) Z1 would immediately
collapse the first bracket of a no-error encoded state to |000i or |111i. In
the next section we thus speak in a more general manner as to how one can
measure syndrome operator eigenvalues without destroying the eigenstate.
But before moving to that issue, let us exemplify how single qubit quan-
tum unitary operations can be done on the encoded states. For this we choose
simply the logical Z and X operators, denoted by ZL and XL respectively.
Exercise .4
Show that for XL = Z1 Z2 Z3 Z4 Z5 Z6 Z7 Z8 Z9 and ZL = X1 X2 X3 X4 X5 X6 X7 X8 X9 ,
As the Shor code correct for only single qubit errors, it is worth consid-
ering the probability p of single qubit errors below which the code will be
useful, i.e., the probability for it to fail to correct is very small. The proba-
bility of no errors is (1 p)9 , while the probability that single qubit errors
occur and are corrected is 9p(1 p)8 . Thus the probability for the code to
fail is (to the second order in a small p)
HA 1
|0iA | Li ! p (|0iA + |1iA ) | Li
2
CZA,2 CZA,1 1
! p (|0iA | Li + |1iA Z1 Z2 | L i)
2
1 1
= p (|0iA | Li + |1iA | L i) = p (|0iA + |1iA ) | Li
2 2
HA
! |0iA | Li for = +1, |1iA | Li for = 1.
6
0 H H 0 A if Z1Z 2 = +1
A
1 A if Z1Z 2 = −1
Z
Z
ψL ψL
7
physical qubit is 3 (X, Z and XZ), implying a total of 3n errors for n qubits.
Additionally there have to be a no error state. All these have to be doubled
up as we have both |0L i and |1L i. All these states have to be accomodable in
the total Hilbert space of n qubits. Thus the minimum number of physical
qubits n required for a code correcting any single qubit error is given by the
criterion
2(3n + 1) 2n .
As you can easily check, the above criterion is not satisfied for either of
n = 2, 3, 4. It is satisfied for n = 5, and indeed a 5 qubit code exists! We
will discuss this next.
It is very easy to remeber the above stabilizer generators as they are cyclic
permutations of XZZZI.
Exercise .5
⇥ ⇤
Show that M (j)2 = 1, M (j) , M (k) = 0. Hints: X 2 = Z 2 = 1 and
XZ = ZX (you will have to use this relation twice in each commutation).
In terms of the stabilizer generators, the logical qubit states are given by
1
|0L i = (1 + M (1) )(1 + M (2) )(1 + M (3) )(1 + M (4) ) |00000i ,
16
1
|1L i = (1 + M (1) )(1 + M (2) )(1 + M (3) )(1 + M (4) ) |11111i .
16
As a rule of thumb, given m stabilizer generators of any stabilizer code
with n physical qubits (which are generally easy to remember), the logical
qubit states are given by
m
1 Y
|0L i = p (1 + M (j) ) |0i⌦n ,
m
2 j=1
8
m
1 Y
|1L i = p (1 + M (j) ) |1i⌦n .
m
2 j=1
Exercise .6
Using the results of the previous exercise and the definition of |0L i and |1L i
show that they are eigenstates of M (j) with eigenvalues +1.
Now, to show the error correction with the 5 qubit code, consider an error
X3 . Now consider the syndrome operator M (1) acting on the state with the
error
Exercise .7
Show that a X1 error can be detected by only M (4) to have 1 eigenvalue,
while Z2 error may be detected by M (3) and M (4) having 1 eigenvalues.
In the same manner as the above exercise, you can show that all possible
single qubit errors can be detected by a combination M (j) attaining negative
eigenvalues.
Exercise .8
Show that XL = X1 X2 X3 X4 X5 and ZL = Z1 Z2 Z3 Z4 Z5 commute with
M (j ). Hence show that XL and ZL are the logical X and Z operators for the 5
qubit code.
9
H H
H H
H H
H H
H H
H H
H H
H H
H H
Z
Z
Z
Z
Z
Z
Z
Z
Z
Figure 3: Transversal circuit for the implementation of a CNOT among two
encoded qubits in the Shor 9 qubit code.
it takes the state in any of the brackets of |1L i of the Shor code takes to a
state with an odd number of 1s (verify this fact yourself).
All fault tolerant quantum computation with the kinds of codes discussed
so far takes place by transversal operations so as to prevent the multiplication
of errors.
10
Quantum Error Correction – Addendum
Sougato Bose
March 17, 2021
1
Encoded Encoded
qubit from qubit to the
earlier part of next part
circuit Syndrome of circuit
Detec:on
& Error
Correc:on
Fault Tolerant
(e.g. Transversal)
Gate
Encoded
Encoded
Encoded qubit to the
qubit from next part
Syndrome
earlier part of of circuit
Detec:on
circuit
& Error
Correc:on
2
concatenation, when we have taken a logical qubit and replaced each of its
physical qubits by a logical qubit, the chance of a single uncorrected error
in each of those logical qubits is Cp2 –so this now replaces the previous p
as the chance of a single qubit error. Thus at this stage the probability
of uncorrected errors is C(Cp2 )2 = (Cp)4 /C. Continuing likewise, in the
kth stage of concatenation, if we obtain the probability of errors remaining
uncorrected to be
k
(Cp)2 /C,
which can be ensured to be small to any desired level – so that a circuit of any
depth (an arbitrarily long quantum circuit) may be carried out for p < 1/C.
Thus the value of the probability pthresh ⇠ 1/C is called the fault tolerant
threshold. We must reduce errors due to decoherence to lower than pThresh in
order to perform arbitrarily long quantum computations by concatenating a
given code a required number of times.
2 Surface Codes
The syndrome measurement operators, as well as the transversal realization
of quantum gates required very long range gates, such as between ancilla
qubit for syndrome measurement and the physical qubit 9 of a logical qubit.
However, it is difficult to have such long range interactions between distant
qubits (interactions between qubits typically fall o↵ with distance). This mo-
tivates a code which is local in terms of its stabilizer generators. An example
of this is the surface code. The surface code has additional advantages as
will be pointed out shortly. A surface code encoding a single logical qubit
is represented in terms of a 2D grid with physical qubits (filled circles) at
the edges as shown in Fig.2. Its stabilizer generators (i.e., syndrome mea-
surement operators) are labelled in terms of the vertex v they surround or
the plaquette p to which they belong (each square of the grid in the figure
supports a plaquette)
Exercise .1
Argue why all the stabilizer generators commute with each other and square
up to the identity. Hints: Again use X 2 = Z 2 = 1 and XZ = ZX for the
first part.
Now we exclude one of the above stabilizers (e.g., the plaquette defined
by grey circles in Fig.2) to define error protected logical states as
nv np
1 Y 1 Y
| Li = p (1 + M (v) ) p (1 + M (p) ) |0i⌦n ,
2nv v=1 2 np p=1,p6=greyed
3
Z
Z Z
Z
X
X
X
X
Figure 2: A surface code encoding a single logical qubit is shown with physical
qubits at the edges. The product of operators X shown around a given vertex,
and the product of Z operators belonging to a given square (these Zs are said
to form a plaquette) define the stabilizer generators.
4
where n is the number of phsyical qubits, np and nv being the number of
plaquettes and vertices respectively. It is straightforward that all the included
operators will have an eigenvalue +1 in the | L i state. As the excluded
plaquette operator can have ±1 eigenvalue, this corresponds to a doubly
degenerate space – there are two distinct states of the form | L i. A single
logical qubit can thus be encoded in these two degenerate states.
It is easy to check that when a X error occurs on a given physical qubit,
both the plaquette stabilizer generators which overlap with it change their
eignevalue to 1, while if a Z error occurs on a given physical qubit then
the two vertex operators which overlap with it change their eigenvalue to
1. Note that the stabilizer generators are all local as they contain only
4 operators of neighbouring qubits. The stabilizer generators can thus be
measured locally by placing syndrome measuring ancillae (called data qubits)
in each vertex and at the centre of each square.
For your knowledge only (not examinable): Two other advantages of the
surface code are that they can give much higher thresholds pthresh ⇠ 0.1 for
fault tolerant quantum computation. Each stabilizer generator term can be
converted to a Hamiltonian term to describe a many-body system where the
ground state is protected by an energy gap. More qubits can be encoded by
suppressing more of the stabilizer generators.
5
Physical Implementations of Quantum
Computation I: Generic Qubit-Oscillator
Hamiltonians
Sougato Bose
March 24, 2021
1
1 Evolution of a quantum system under per-
turbations: the interaction picture and the
Dyson series
Consider the evolution of a quantum system under a Hamiltonian
H = H0 + V (t),
where H0 is a time-independent dominant part of the Hamiltonian and V (t)
is a perturbation, which may be time dependent. The ket in the standard
Schroedinger equation, which we will label as | S (t)i then follows
@| S (t)i
i = H | S (t)i ,
@t
where frequency units have been used for H (we have taken h̄ = 1). If we
define the interaction picture ket by
| I (t)i = eiH0 t | S (t)i ,
We will use the above first 3 terms for identifying e↵ective time-independent
Hamiltonians under which systems evolve. That will help us understand
how qubits are controlled and coupled by classical and quantum harmonic
oscillators respectively. Why does it suffice to calculate evolutions in the
interaction picture? In quantum mechanics, we are, in the end, interested
in computing correctly the amplitude of various states |mS i i.e, on hmS | S i.
But hmS | S i = hmI | I i, so that computing the evolution entirely in the
interaction picture suffices.
3
⌦ e i t 1 ei t 1
⇡1 i {e i + ei + } (As !0 >> ⌦).
4 i i
Now let us choose ! ⇡ !0 (this is called resonance) so that = |!0 !| ! 0.
Thus t << 1/ for a substantially long time t, during which the above
unitary evolution reduces to
⌦ i
UI (t) = 1 i {e + ei + }t
4
⌦
=1
i (cos X sin Y )t.
2
The above unitary evolution in the interaction picture could thus be regarded
as being generated by the e↵ective Hamiltonian
⌦
VRot = (cos X sin Y ).
2
This Hamiltonian is called a rotating wave approximation of the interaction
picture Hamiltonian VI (t) and often heuristically justified as obtained by
neglecting the fast rotating terms in VI (t). The derivation presented here
using the Dyson series gives a more rigorous justification. It also justifies the
picture of resonance that you might know from before, namely that the right
frequency of a field !, only when it matches the qubit energy separation in
frequency units (!0 ), it drives a transition between the |0i and |1i states in
the strongest possible manner.
It is clear that by choosing (i) resonance, and (ii) the appropriate ,
VRot can be used to implement any rotation in the Bloch-sphere about any
axis in the XY plane, including both X and Y gates and arbitrary rotations
about these axes. Thus arbitrary unitary operations on the qubit are already
possible (remember, that 3 operations about 2 axes in the Bloch sphere suffice
for any SU (2) operation). However, there is also a di↵erent way of performing
the Z operation directly by using a large detuning field satisfying >> ⌦,
but still << !, !0 . In that case, as you can see from the above, the second
term in the Dyson series goes to zero. Thus the third term is the important
one. The unitary evolution thus becomes (setting = 0, and neglecting
⌦/(! + !0 ) << 1 terms)
Z t i t0
⌦ i!t0 i!t0 i!0 t0 i!0 t0 e 1
UI (t) = 1 ( )2 { (e +e )( +e + e )(
4 0 i
t0
ei 1 0
+ + )dt }
i
Z t i t0 i t0
⌦ t0 e 1 i t0 e 1
=1 ( )2 { + e i
+ +e }dt0
4 0 i i
2
⌦ ⌦
=1 i( )2 [ +, ]t (Neglecting 2
terms)
4
⌦2
=1 i Zt.
4
4
The above unitary is induced by an e↵ective Hamitonian
⌦2
VDet = Z.
4
Thus a Z gate and arbitrary Z rotations of a qubit are possible by ap-
plying, for an appropriate time t, the e↵ective Hamiltonian VDet which arises
by choosing ⌦ << << !, !0 .
The above e↵ective Hamiltonian also gives the condition for no quantum
gate on a given qubit even when a classical harmonic oscillator interacts with
it with a strength ⌦. The condition is to choose such a large detuning
2
such that even this third term of the Dyson series is zero, i.e., ⌦ t << 1
so that for times t satisfying this, essentially no evolution is obtained (the
fourth term onwards in the Dyson series will have even lower contribution
to the evolution than the third term). This condition is needed if there are
several qubits on which a single classical field is incident, but we want only
one of them to undergo a quantum gate. We can then very far detune the
frequency of the field from the energy separations between the states of all
qubits but the given one. On the given one, we fix and appropriately to
have any single qubit gate.
Exercise .1
Repeat the calculations of the previous section with H0 = !20 Z and VI =
g(ae i!t + a† ei!t )X for = ! !0 ⇡ 0 and using g << !, !0 , to get, from
the second term of the Dyson series, an e↵ective Hamiltonian
VJCM = g(a + + a† ).
Once you see the above form of VJCM , which is called the Jaynes-Cummings
Hamiltonian, it is easy to interpret it as one inducing a flip-flop process of
energy between the qubit and the oscillator. As the frequencies of the oscil-
lator and the qubit are matched (! ⇡ !0 ), a quanta of energy !0 goes from
oscillator to qubit via a + term while a quanta of energy !0 goes from qubit
to oscillator via a† term.
Now consider the case of is nonzero but still << !, !0 . If you cleverly do
not complete the integration over t0 in the second term of the Dyson series
0
for the terms with e±i t dependence (assuming they would be finite and
remain relevant), but performs the integration over the terms e±i(!+!0 )t and
subsequently neglect these terms as g << !, !0 , you will see that the unitary
evolution can be regarded as being generated by the e↵ective Hamiltonian
i t †
VJCM-Det = g(ei t a + +e a ).
6
P θ (t) = (ω + χ )t for 0 Q
mω m !
δp ~
2
α θ (t) = (ω − χ )t for 1
Q
θ
δx
X
!
~
2mω m
Figure 2: A coherent state and its evolution due to the dispersive interaction
with a qubit.
where |ni are the di↵erent number states of the harmonic oscillator. Coher-
ent states of oscillators are most easily prepared in the laboratory and have
a representation in phase space as a circle about a centre x = 2 x Re(↵), p =
2 p Im(↵) as shown in Fig.2. The ↵coherent state of an isolated oscillator
evolves in time uniformly as |↵|e✓(t) with ✓(t) = !t due to the action of its
own Hamiltonian !n̂. However, as can be understood by applying the Hamil-
tonian VDisp to the coherent state, the rate of evolution of ✓(t) is a↵ected by
the state of the qubit. For the state |0iQ of the qubit it is ✓(t) = (! + )t,
while for the qubit in the state |1iQ it is ✓(t) = (! )t (we have introduced
the qubit label Q here to di↵erentiate from oscillator number states). Thus,
after some time measuring the position/momentum of the oscillator, one can
deduce whether the qubit was in the state |0iQ or |1iQ . Such a measurement
is called a quantum non-demolition measurment as the computational basis
states of the qubit are una↵ected. In case when a qubit is coupled to a quan-
tum electromagnetic field oscillator, the position/momentum get replaced by
electric/magnetic fields (these measurements are equivalent to determining
how much the phase of the field has changed). So the appropriate variable
of the field can be measured to determine the ✓ and thereby the qubit state.
This method is used to measure the state of a superconducting qubit by
7
allowing it to shift the phase of a quantum microwave field coupled to it.
Z t Z t0
(1) (2) (1) 00 (2) 00 00 0
UI (t) = 1 (VJCM-Det (t0 )+VJCM-Det (t0 )) (VJCM-Det (t )+VJCM-Det (t ))dt dt
0 0
The two qubit interaction part is thus described by the e↵ective Hamiltonian
Exercise .2
Starting from a state 01 of two qubits show that time evolution due to
VFlip-flop can be used to maximally entangle two qubits. Write down the
full unitary operation corresponding to the gatepat this ideal time when it
entangles the qubits maximally (this is called a iSWAP gate).
The above method is used to perform a two qubit gate on superconducting
qubits coupled by a microwave cavity bus – this enables keeping the qubits
at some separation and gives more flexibility in wiring so that diverse archi-
tectures of connected qubits can be generated! You 2D wire a set of qubits
through microwave cavities which can be thin strips of metal in arbitrary
shapes. Of course, superconducting qubits can be made to interact directly
as well when placed adjacently and coupled through the charges of the qubit
states (Capacitive coupling) or the fluxes of the qubit states (Inductive cou-
pling).
There is another type of two qubit gate which can be generated when the
qubit levels |0i and |1i have an energy separation !0 << g. In this case, the
X operator in interaction picture hardly have a time-dependence and thereby
remains as X, while the oscillator operators have all the time dependence so
that
(j)
VI (t) = g(ae i!t
+ a† ei!t )X (j) .
If ! >> g the second term in the Dyson series can again become negligible,
so that we concentrate on the third term to get
8
Z t Z t0
(1) (2) (1) 00 (2) 00 00 0
UI (t) = 1 (VI (t0 ) + VI (t0 )) (VI (t ) + VI (t ))dt dt
0 0
g 2 (1) (2)
= 1+i X X t (omitting constant terms),
!
so that an e↵ective Hamiltonian is
g 2 (1) (2)
VMS = X X .
!
We know that the above is identical in form to the Ising Hamiltoninan /
Z (1) Z (2) studied earlier, so up to a basis change, it can also be used to perform
a CZ gate between two qubits. The above type of gate is called the Moelmer-
Sorensen gate. It is used for ion trap qubits where the qubits are encoded in
ionic levels and a common mode of vibration of the ions acts as the quantum
haronic oscillator. This gate has the advantage (as you can see above) that
its strength is independent of the quantum state of the mediating oscillator –
no variable of the mediating oscillator is present in VMS . Thus it is incredibly
robust to thermal heating of the oscillator mode.
In the next part of the lecture, we are going to discuss in more detail the
ion-trap implementation of quantum computation.
9
Physical Implementations of Quantum
Computation II: General Criteria and Ion
Trap Implementation
Sougato Bose
March 25, 2021
1 DiVincenzo Criteria
I label the criteria by letters, rather than numbers, so that they are easier to
remember:
Q: Qubits We need well defined qubits. What does this mean? A valid
qubit can be made by using any two orthogonal states of a quantum system
as |0i and |1i as long as we can stay within the subspace defined by these
two sates during our entire computation. For example, a spin-1/2 system is
a natural qubit – no matter what we do to it, we will never get beyond the
states defined by superpositions of |#i = |0i and |"i = |1i. However, we may
sometimes want to use higher dimensional systems, indeed even continuous
variable systems, as qubits. For example, atoms. With the right frequency
and polarization it is possible to choose to work solely in the subspace de-
fined by a certain pair of atomic states |gi = |0i (g standing for ground)
and |ei = |1i (e standing for excited) while having a vanishing/insignificant
amplitude of involving any other states of the atom. The number states of a
harmonic oscillator, on the other hand, cannot be used a qubit, as any field
which induces a transition between number states |0i and |1i also induces
a transition between |1i and |2i. However, an anharmoic oscillator can be
used as a qubit, with the classic example being superconducting transmon
qubits, where the anharmonicity enables the states |0i and |1i to addressed
by a di↵erent frequency than the states |1i and |2i.
This criteria can obviously be somewhat relaxed in the sense that one
can use qudits defined in any system and use multiple qudits (again defined
1
according to the criterion that one should not go out of the subspace defined
by the d states of the qudit). One can even directly use an exponentially
large Hilbert space of any system even if there are no “localized” qubits is
them. The typical example of that is topological quantum computing which
is done by braiding anyons, but the anyons themselves are do not encode
qubits locally in them. The global state evolves in an exponentially (in the
number of anyons) large subspace when the anyons are braided.
2
Quantum processors in the near future will be
small!
--- bus mode, fabrication, efficiency
Qubits/Entanglement
Usually the use of photons are envisaged for making the links
Figure 1: A larger quantum computer being formed from several smaller
quantum processors, with common bus of each processor shown as blue thin
rectangles. Qubits or entanglement needs to flow among the processors to
make it a larger multi-processor quantum computer.
S: Scaling One must be able to enlarge the Hilbert space present for
processing to a large amount. Typically it means the ability to increase the
number of qubits taking part in a quantum computation. Typically each
individual processor of a quantum computer is limited in size in terms of
numbers of qubits. This is because they often work by means of a common
bus (as in the case of ion-trap quantum computing) which gets more sluggish
when there are more ions connected to the same bus. Alternatively quantum
computers some times work through local interactions between neighbouring
qubits (such as spin-exchange based, which happens only between proximal
spins). Then how to do long range gates remains open. Thus it seems that we
need to network multiple quantum processors together to make an e↵ective
larger quantum computer as shown in Fig.1. This makes it necessary to be
able to do gates between qubits in one register and those of another register.
For that one either needs to transfer qubits between these processors or needs
to share nearly perfect entangled states (e.g., Bell-states) between them so
that qubits can be moved by teleportation. In order to move, the usually
static qubits of one register must be transformed to mobile qubits – thus this
criteria of scaling is often stated as the ability to interconvert between static
and flying qubits.
3
One can remember the DiVincenzo criteria in terms of the acronym QIPONS,
which in turn can be remembered through the phrase quantum information
processing overcoming noise sources (note that certain words of this phrase,
particularly overcoming and sources do not correspond to their meaning in
QIPONS).
Q: Two of the chosen internal levels of each ion is used as a qubit. These
are typically chosen, as shown in Fig.3, as two hyperfine levels labelled as |ei
and |gi respectively of the ground state manifold – there are no other states
below to which these states can spontaneously decay through purely a single
photon emmision, so that these are stable on their own (without the action
of external fields). When all fields except ⌦1 and ⌦2 are switched o↵, then
the qubit’s evolution remains within the subspace spanned by |ei and |gi, as
for a large detuning , the intermediate level |ri does not get populated (see
the point P on processing for further clarity on this point) so that we have
an ideal two level system.
4
Figure 2: The architecture of a linear ion trap, showing the cross section
fields as well as surface trap versions.
5
r a2
a3 Δ
a1
Ωf
Ω2
Ω1
δ
e
g
Figure 3: The level configuration for an ion trap qubit and all the transitions
between them that are necessary for initialization, processing and readout
(for clarity more levels have been shown than necessary; some levels can be
used for dual purpose). The levels |ei and |gi serve as the relevant qubit
levels.
6
I: The initialization is done through a process called optical pumping. In this
process, we assume that at first the atom is with equal probabilities in both
the states |ei and |gi, both being states of the ground state manifold (after
say all spontaneous decays have taken place). Now two lasers of appropriate
frequencies and polarizations are applied to the atom for the |ei level to go
to the level |a2 i via the level |a1 i as shown in Fig. 3. The level |a2 i, in
turn, selectively spontaneously decays via a single photon transition to the
level |gi (wiggly line in the figure). It cannot decay spontaneously to any
other level |g 0 i, say. However, once in |gi, no extra processes are happening
to the atom – thus the probability starts collecting here – it is called a dark
state. Thus all probability in |ei can be converted to a probability of |gi.
The longer duration for which the two lasers doing the |ei ! |a2 i transition
are kept on, higher the certainty of intialization to |gi, which can serve as
the |0i qubit state. In principle, the motional states of the collection of ions
should also be cooled as much down as possible – this can be accomplished
by a procedure called side-band cooling, which will become clearer in point P
below (though this cooling is not pivotal; as we have seen in the previous part
of this lecture, the Moelmer-Sorensen gates are independent of the cooling).
with ⌦1 ⌦2 , which can be used for qubit manipulation (we have introduced
the subscript Q for qubit). By introducing relative phases of the lasers, as in
7
the earlier part of this lecture, any rotation axis in the equator of the Bloch
sphere is possible, and by using only one of the lasers (say, by inducing just
⌦1 ) and with >> ⌦1 a rotation about the z axis is also possible. Thus
arbitrary single qubit gates are possible.
What about two-qubit gates? For that we need to consider the Hamilto-
nian including the spatial profile of the Electric field of the laser. The spatial
profile (for which we expand eikx = 1 + ikx + ..., where k = 2⇡/ is the wave
vector of the laser along the trap axis and x is the position of the ion for the
case in which is confined well within the wavelength – this is called the
Lamb-Dicke regime) gives a position dependent coupling strength between
the ion’s qubit states; if the position is then quantized then e↵ectively the
ion’s position, and through its Coulomb interactions with other ions, the
common vibrational modes of all the ions, are the quantum harmonic os-
cillators which couple to the qubit. Considering just the common centre of
mass mode of all the ions (all the modes are uncoupled to each other under
the harmonic approximation), the position of the ion can be taken as the
variable denoting the position coordinate of the centre of mass of all ions,
and accordingly quantized as
x̂ = x (a + a† ),
q
h̄
with x = 2m! m
where m is the collective mass, and !m the centre of mass
frequency, of all the ions. The Hamiltonian in the interaction picture, using
the fact that now we use 6= 0, so that it is the e↵ective detuning of the
qubit transition (|ei hg| = + , |gi he| = ), and quantizing, is
VIon-JCM = ⌦⇤ ⌘(a + + a† ),
VIon-AJCM = ⌦⇤ ⌘(a + + a† ),
which can be called the anti-JCM model. Note that the JCM Hamiltonian
annihilates a phonon (quanta of energy of the motion) from the common
motional mode everytime the atom goes from |gi to |ei. This process can
be exploited to cool the common motional mode through a process called
sideband cooling. Just as discussed in the pumping based atomic state ini-
tialization, the two lasers performing the |ei ! |a2 i transition are switched
8
on, so that each time the atom performs a |gi to |ei transition and thereby
grabs a phonon, it is immediately sent to |a2 i from which it comes back to
|gi by spontaneous emission, so the process can be repeated again, to grab
more phonons and cool the centre of mass motional state further. Now note
that if we switch on both VIon-JCM and VIon-AJCM simultaneously (that will
require four lasers), we will have
VIon-MS = ⌦⇤ ⌘X(a + a† ),
which we know, can be used with two ions i and j to construct a Molmer
Sorensen two qubit gate equivalent to the action of a X basis Ising Hamil-
tonian of the form Xi Xj for a certain interval of time [Note: The gate can
also be done in the non-perturbative limit using VIon-MS of two ions alone,
in which the motional mode entangles and subsequently disentangles from
the qubits after one oscillation period 2⇡/!m , which is what is traditionally
called the Moelmer Sorensen gate].
O: For measuring the trapped ion qubits states, we use the phenomenon
of resonance fluorescence. Consider no laser applied, and we only want to
measure whether the state of the atom is |ei of |gi. We apply a laser to the
atom so that a transition from |gi to |a3 i takes place (strength ⌦f ), which
immediately spontaneously decays down to |gi emitting a photon (wiggly
line in Fig.3). This will again be excited by the applied laser to emit again.
In this way a lot of photons come out of the atom in all directions if it was
initially in the |gi state (it fluoresces), while it remains dark if it was in the
|ei state as no lasers are applied to it. After collecting n photons in directions
di↵erent from the original laser direction, your uncertainty that a fluorescing
atom was in the state |gi is reduced by 1/sqrtn (shot noise limit), so that
by observing for a required amount of time you can reduce your inaccuracy
of your state measurement as much as you require.
9
Fig.4 which shows the shuttling of data qubit ions to bring close to measure
qubit ions.
S: A single line of ions cannot be made arbitrarily long because the couplings
are / 1/sqrtm (see point P) where m is the total mass of ions. Thus it is
envisaged that multiple registers of several tens of ions, each a linear one,
as shown as the horizontal zones in Fig.4. These can then be connected by
“shuttling” ions, where an ion from one of these registers is taken and phys-
ically moved, by changing voltages among successive flakes of the electrodes
(each linear electrode is made of several flakes). This still su↵ers from being
a sluggish method. The other approach is to use entanglement swapping: we
generate ion-photon entangled states in each register and jointly detect the
photons in a Bell state to entangle the ionic qubits. This also su↵er from low
entanglement generation efficiency, partly to do with the collection efficiency
of photons (imporvable by incorporating cavities).
This ends our description of ion trap quantum computation as one of the
examples of the physical implementations of quantum computation. I hope
that the challenges are clear to you.
10
Taken from: https://advances.sciencemag.org/content/3/2/e1601540
Figure 4: Figure showing both the surface code realization in ion trap arrays,
as well as multiple registers being connected by the shuttling of ions in 2D
surface traps.
11