Compilation#

So far, we have already covered enough to be able to design the Circuit s we want to run, submit them to a Backend, and interpret the results in a meaningful way. This is all you need if you want to just try out a quantum computer, run some toy examples and observe some basic results. We actually glossed over a key step in this process by using the get_compiled_circuit() method. The compilation step maps from the universal computer abstraction presented at Circuit construction to the restricted fragment supported by the target Backend, and knowing what a compiler can do to your program can help reduce the burden of design and improve performance on real devices.

The necessity of compilation maps over from the world of classical computation: it is much easier to design correct programs when working with higher-level constructions that aren’t natively supported, and it shouldn’t require a programmer to be an expert in the exact device architecture to achieve good performance. There are many possible low-level implementations on the device for each high-level program, which vary in the time and resources taken to execute. However, because QPUs are analog devices, the implementation can have a massive impact on the quality of the final outcomes as a result of changing how susceptible the system is to noise. Using a good compiler and choosing the methods appropriately can automatically find a better low-level implementation. Each aspect of the compilation procedure is exposed through pytket to provide users with a way to have full control over what is applied and how.

The primary goals of compilation are two-fold: solving the constraints of the Backend to get from the abstract model to something runnable, and optimising/simplifying the Circuit to make it faster, smaller, and less prone to noise. Every step in compilation can generally be split up into one of these two categories (though even the constraint solving steps could have multiple solutions over which we could optimise for noise).

Each compiler pass inherits from the BasePass class, capturing a method of transforming a Circuit. The main functionality is built into the BasePass.apply() method, which applies the transformation to a Circuit in-place. The get_compiled_circuit() method is a wrapper around the apply() from the Backend ‘s recommended pass sequence. This chapter will explore these compiler passes, the different kinds of constraints they are used to solve and optimisations they apply, to help you identify which ones are appropriate for a given task.

Predicates#

Solving the constraints of the target Backend is the essential goal of compilation, so our choice of passes is mostly driven by this set of constraints. We already saw in the last chapter that the required_predicates property gives a collection of Predicate s, describing the necessary properties a Circuit must satisfy in order to be run.

Each Predicate can be constructed on its own to impose tests on Circuit s during construction.

from pytket import Circuit, OpType
from pytket.predicates import GateSetPredicate, NoMidMeasurePredicate

circ = Circuit(2, 2)
circ.Rx(0.2, 0).CX(0, 1).Rz(-0.7, 1).measure_all()

gateset = GateSetPredicate({OpType.Rx, OpType.CX, OpType.Rz, OpType.Measure})
midmeasure = NoMidMeasurePredicate()

print(gateset.verify(circ))
print(midmeasure.verify(circ))

circ.S(0)

print(gateset.verify(circ))
print(midmeasure.verify(circ))
True
True
False
False

Common Predicate

Constraint

GateSetPredicate

Every gate is within a set of allowed OpType s

ConnectivityPredicate

Every multi-qubit gate acts on adjacent qubits according to some connectivity graph

DirectednessPredicate

Extends ConnectivityPredicate where OpType.CX gates are only supported in a specific orientation between adjacent qubits

NoClassicalControlPredicate

The Circuit does not contain any gates that act conditionally on classical data

NoMidMeasurePredicate

All OpType.Measure gates act at the end of the Circuit (there are no subsequent gates on either the Qubit measured or the Bit written to)

When applying passes, you may find that you apply some constraint-solving pass to satisfy a particular Predicate, but then a subsequent pass will invalidate it by, for example, introducing gates of different gate types or changing which qubits interact via multi-qubit gates. To help understand and manage this, each pass has a set of pre-conditions that specify the requirements assumed on the Circuit in order for the pass to successfully be applied, and a set of post-conditions that specify which Predicate s are guaranteed to hold for the outputs and which are invalidated or preserved by the pass. These can be viewed in the API reference for each pass.

Users may find it desirable to enforce their own constraints upon circuits they are working with. It is possible to construct a UserDefinedPredicate in pytket based on a function that returns a True/False value.

Below is a minimal example where we construct a predicate which checks if our Circuit contains fewer than 3 CX gates.

from pytket.circuit import Circuit, OpType
from pytket.predicates import UserDefinedPredicate

def max_cx_count(circ: Circuit) -> bool:
    return circ.n_gates_of_type(OpType.CX) < 3

# Now construct our predicate using the function defined above
my_predicate = UserDefinedPredicate(max_cx_count)

test_circ = Circuit(2).CX(0, 1).Rz(0.25, 1).CX(0, 1) # Define a test Circuit

my_predicate.verify(test_circ)
# test_circ satisfies predicate as it contains only 2 CX gates
True

Rebases#

One of the simplest constraints to solve for is the GateSetPredicate, since we can just substitute each gate in a Circuit with an equivalent sequence of gates in the target gateset according to some known gate decompositions. In pytket, such passes are referred to as “rebases”. The intention here is to perform this translation naively, leaving the optimisation of gate sequences to other passes. Rebases can be applied to any Circuit and will preserve every structural Predicate, only changing the types of gates used.

from pytket import Circuit
from pytket.passes import RebaseTket

circ = Circuit(2, 2)
circ.Rx(0.3, 0).Ry(-0.9, 1).CZ(0, 1).S(0).CX(1, 0).measure_all()

RebaseTket().apply(circ)

print(circ.get_commands())
[TK1(0, 0.3, 0) q[0];, TK1(0.5, 3.1, 3.5) q[1];, TK1(0.5, 0.5, 0.5) q[1];, CX q[0], q[1];, TK1(0, 0, 0.5) q[0];, TK1(0.5, 0.5, 0.5) q[1];, CX q[1], q[0];, Measure q[0] --> c[0];, Measure q[1] --> c[1];]

RebaseTket is a standard rebase pass that converts to CX and TK1 gates. This is the preferred internal gateset for many pytket compiler passes. However, it is possible to define a rebase for an arbitrary gateset. Using RebaseCustom, we can provide an arbitrary set of one- and two-qubit gates. Rather than requiring custom decompositions to be provided for every gate type, it is sufficient to just give them for OpType.CX and OpType.TK1. For any gate in a given Circuit, it is either already in the target gateset, or we can use known decompositions to obtain a OpType.CX and OpType.TK1 representation and then map this to the target gateset.

from pytket import Circuit, OpType
from pytket.passes import RebaseCustom

gates = {OpType.Rz, OpType.Ry, OpType.CY, OpType.ZZPhase}
cx_in_cy = Circuit(2)
cx_in_cy.Rz(0.5, 1).CY(0, 1).Rz(-0.5, 1)

def tk1_to_rzry(a, b, c):
    circ = Circuit(1)
    circ.Rz(c + 0.5, 0).Ry(b, 0).Rz(a - 0.5, 0)
    return circ

custom = RebaseCustom(gates, cx_in_cy, tk1_to_rzry)

circ = Circuit(3)
circ.X(0).CX(0, 1).Ry(0.2, 1)
circ.ZZPhase(-0.83, 2, 1).Rx(0.6, 2)

custom.apply(circ)

print(circ.get_commands())
[Rz(0.5) q[0];, Rz(0.5) q[1];, Ry(1) q[0];, Rz(3.5) q[0];, CY q[0], q[1];, Rz(3.5) q[1];, Ry(0.2) q[1];, ZZPhase(3.17) q[2], q[1];, Rz(0.5) q[2];, Ry(0.6) q[2];, Rz(3.5) q[2];]

For some gatesets, it is not even necessary to specify the CX and TK1 decompositions: there is a useful function auto_rebase_pass() which can take care of this for you. The pass returned is constructed from the gateset alone. (It relies on some known decompositions, and will raise an exception if no suitable known decompositions exist.) An example is given in the “Combinators” section below.

A similar pair of methods, SquashCustom() and auto_squash_pass(), may be used to construct a pass that squashes sequences of single-qubit gates from a given set of single-qubit gates to as short a sequence as possible. Both take a gateset as an argument. SquashCustom() also takes a function for converting the parameters of a TK1 gate to the target gate set. (Internally, the compiler squashes all gates to TK1 and then applies the supplied function.) auto_squash_pass() attempts to do the squash using known internal decompositions (but may fail for some gatesets). For example:

from pytket.circuit import Circuit, OpType
from pytket.passes import auto_squash_pass

gates = {OpType.PhasedX, OpType.Rz, OpType.Rx, OpType.Ry}
custom = auto_squash_pass(gates)

circ = Circuit(1).H(0).Ry(0.5, 0).Rx(-0.5, 0).Rz(1.5, 0).Ry(0.5, 0).H(0)
custom.apply(circ)
print(circ.get_commands())
[PhasedX(2.5, 0) q[0];, Rz(0.5) q[0];]

Note that the H gates (which are not in the specified gateset) are left alone.

Placement#

Initially, a Circuit designed without a target device in mind will be expressed in terms of actions on a set of “logical qubits” - those with semantic meaning to the computation. A placement (or initial mapping) is a map from these logical qubits to the physical qubits of the device that will be used to carry them. A given placement may be preferred over another if the connectivity of the physical qubits better matches the interactions between the logical qubits caused by multi-qubit gates, or if the selection of physical qubits has better noise characteristics. All of the information for connectivity and noise characteristics of a given Backend is wrapped up in a BackendInfo object by the Backend.backend_info property.

The placement only specifies where the logical qubits will be at the start of execution, which is not necessarily where they will end up on termination. Other compiler passes may choose to permute the qubits in the middle of a Circuit to either exploit further optimisations or enable interactions between logical qubits that were not assigned to adjacent physical qubits.

A placement pass will act in place on a Circuit by renaming the qubits from their logical names (the UnitID s used at circuit construction) to their physical addresses (the UnitID s recognised by the Backend). Classical data is never renamed.

from pytket import Circuit
from pytket.extensions.qiskit import IBMQBackend
from pytket.passes import PlacementPass
from pytket.predicates import ConnectivityPredicate
from pytket.placement import GraphPlacement

circ = Circuit(4, 4)
circ.H(0).H(1).H(2).V(3)
circ.CX(0, 1).CX(1, 2).CX(2, 3)
circ.Rz(-0.37, 3)
circ.CX(2, 3).CX(1, 2).CX(0, 1)
circ.H(0).H(1).H(2).Vdg(3)
circ.measure_all()

backend = IBMQBackend("ibmq_quito")
place = PlacementPass(GraphPlacement(backend.backend_info.architecture))
place.apply(circ)

print(circ.get_commands())
print(ConnectivityPredicate(backend.backend_info.architecture).verify(circ))
[H node[0];, H node[1];, H node[3];, V node[4];, CX node[0], node[1];, CX node[1], node[3];, CX node[3], node[4];, Rz(3.63*PI) node[4];, CX node[3], node[4];, CX node[1], node[3];, Vdg node[4];, Measure node[4] --> c[3];, CX node[0], node[1];, H node[3];, Measure node[3] --> c[2];, H node[0];, H node[1];, Measure node[0] --> c[0];, Measure node[1] --> c[1];]
True

In this example, the placement was able to find an exact match for the connectivity onto the device.

In some circumstances, the best location is not fully determined immediately and is deferred until later in compilation. This gives rise to a partial placement (the map from logical qubits to physical qubits is a partial function, where undefined qubits are renamed into an unplaced register).

from pytket import Circuit
from pytket.extensions.qiskit import IBMQBackend
from pytket.passes import PlacementPass
from pytket.placement import LinePlacement

circ = Circuit(4)
circ.CX(0, 1).CX(0, 2).CX(1, 2).CX(3, 2).CX(0, 3)

backend = IBMQBackend("ibmq_quito")
place = PlacementPass(LinePlacement(backend.backend_info.architecture))
place.apply(circ)

print(circ.get_commands())
[CX node[2], node[1];, CX node[2], node[3];, CX node[1], node[3];, CX unplaced[0], node[3];, CX node[2], unplaced[0];]

A custom (partial) placement can be applied by providing the appropriate qubit map.

from pytket.circuit import Circuit, Qubit, Node
from pytket.placement import Placement

circ = Circuit(4)
circ.CX(0, 1).CX(0, 2).CX(1, 2).CX(3, 2).CX(0, 3)

q_map = {Qubit(0) : Node(3), Qubit(2) : Node(1)}
Placement.place_with_map(circ, q_map)

print(circ.get_commands())
[CX node[3], q[1];, CX node[3], node[1];, CX q[1], node[1];, CX q[3], node[1];, CX node[3], q[3];]

A custom placement may also be defined as a pass (which can then be combined with others to construct a more complex pass).

from pytket.circuit import Circuit, Qubit, Node
from pytket.passes import RenameQubitsPass

circ = Circuit(4)
circ.CX(0, 1).CX(0, 2).CX(1, 2).CX(3, 2).CX(0, 3)

q_map = {Qubit(0) : Qubit("z", 0), Qubit(2) : Qubit("z", 1)}
rename = RenameQubitsPass(q_map)
rename.apply(circ)

print(circ.get_commands())
[CX z[0], q[1];, CX z[0], z[1];, CX q[1], z[1];, CX q[3], z[1];, CX z[0], q[3];]

Several heuristics have been implemented for identifying candidate placements. For example, LinePlacement will try to identify long paths on the connectivity graph which could be treated as a linear nearest-neighbour system. GraphPlacement will try to identify a subgraph isomorphism between the graph of interacting logical qubits (up to some depth into the Circuit) and the connectivity graph of the physical qubits. Then NoiseAwarePlacement extends this to break ties in equivalently good graph maps by looking at the error rates of the physical qubits and their couplers. The latter two can be configured using e.g. modify_config() to change parameters like how far into the Circuit it will look for interacting qubits (trading off time spent searching for the chance to find a better placement).

from pytket import Circuit
from pytket.extensions.qiskit import IBMQBackend
from pytket.passes import PlacementPass
from pytket.predicates import ConnectivityPredicate
from pytket.placement import GraphPlacement

circ = Circuit(5)
circ.CX(0, 1).CX(1, 2).CX(3, 4)
circ.CX(0, 1).CX(1, 2).CX(3, 4)
circ.CX(0, 1).CX(1, 2).CX(3, 4)
circ.CX(0, 1).CX(1, 2).CX(3, 4)
circ.CX(0, 1).CX(1, 2).CX(3, 4)
circ.CX(1, 4)   # Extra interaction hidden at higher depth than cutoff

backend = IBMQBackend("ibmq_quito")
g_pl = GraphPlacement(backend.backend_info.architecture)
connected = ConnectivityPredicate(backend.backend_info.architecture)

PlacementPass(g_pl).apply(circ)
print(connected.verify(circ))   # Imperfect placement because the final CX was not considered

# Default depth limit is 5, but there is a new interaction at depth 11
g_pl.modify_config(depth_limit=11)

PlacementPass(g_pl).apply(circ)
print(connected.verify(circ))   # Now have an exact placement
False
True

Mapping#

The heterogeneity of quantum architectures and limited connectivity of their qubits impose the strict restriction that multi-qubit gates are only allowed between specific pairs of qubits. Given it is far easier to program a high-level operation which is semantically correct and meaningful when assuming full connectivity, a compiler will have to solve this constraint. In general, there won’t be an exact subgraph isomorphism between the graph of interacting logical qubits and the connected physical qubits, so this cannot be solved with placement alone.

One solution here, is to scan through the Circuit looking for invalid interactions. Each of these can be solved by either moving the qubits around on the architecture by adding OpType.SWAP gates until they are in adjacent locations, or performing a distributed entangling operation using the intervening qubits (such as the “bridged-CX” OpType.BRIDGE which uses 4 CX gates and a single shared neighbour). The routing procedure used in the pytket RoutingPass takes a placed Circuit and inserts gates to reduce non-local operations to sequences of valid local ones.

from pytket import Circuit
from pytket.extensions.qiskit import IBMQBackend
from pytket.passes import PlacementPass, RoutingPass
from pytket.placement import GraphPlacement

circ = Circuit(4)
circ.CX(0, 1).CX(0, 2).CX(1, 2).CX(3, 2).CX(0, 3)
backend = IBMQBackend("ibmq_quito")
PlacementPass(GraphPlacement(backend.backend_info.architecture)).apply(circ)
print(circ.get_commands())  # One qubit still unplaced
                            # node[0] and node[2] are not adjacent

RoutingPass(backend.backend_info.architecture).apply(circ)
print(circ.get_commands())
[CX node[1], node[0];, CX node[1], node[2];, CX node[0], node[2];, CX unplaced[0], node[2];, CX node[1], unplaced[0];]
[CX node[1], node[0];, CX node[1], node[2];, SWAP node[0], node[1];, CX node[1], node[2];, SWAP node[1], node[3];, CX node[1], node[2];, CX node[0], node[1];]

As shown here, if a partial placement is used, the routing procedure will allocate the remaining qubits dynamically. We also see that the logical qubits are mapped to different physical qubits at the start and end because of the inserted OpType.SWAP gates, such as q[1] starting at node[0] and ending at node[3].

RoutingPass only provides the default option for mapping to physical circuits. The decision making used in Routing is defined in the RoutgMethod class and the choice of RoutingMethod used can be defined in the FullMappingPass compiler pass for producing physical circuits.

Decomposing Structures#

The numerous Box structures in pytket provide practical abstractions for high-level operations to assist in Circuit construction, but need to be mapped to low-level gates before we can run the Circuit. The DecomposeBoxes pass will unwrap any CircBox, substituting it for the corresponding Circuit, and decompose others like the Unitary1qBox and PauliExpBox into efficient templated patterns of gates.

from pytket.circuit import Circuit, CircBox, PauliExpBox
from pytket.passes import DecomposeBoxes
from pytket.pauli import Pauli

sub = Circuit(2)
sub.CZ(0, 1).T(0).Tdg(1)
sub_box = CircBox(sub)
circ = Circuit(4)
circ.Rx(0.42, 2).CX(2, 0)
circ.add_gate(sub_box, [0, 1])
circ.add_gate(sub_box, [2, 3])
circ.add_gate(PauliExpBox([Pauli.X, Pauli.Y, Pauli.Y, Pauli.Y], 0.2), [0, 1, 2, 3])

DecomposeBoxes().apply(circ)
print(circ.get_commands())
[Rx(0.42) q[2];, CX q[2], q[0];, CZ q[0], q[1];, CZ q[2], q[3];, T q[0];, Tdg q[1];, T q[2];, Tdg q[3];, H q[0];, V q[1];, V q[2];, V q[3];, CX q[1], q[0];, CX q[3], q[2];, CX q[2], q[0];, Rz(0.2) q[0];, CX q[2], q[0];, CX q[1], q[0];, CX q[3], q[2];, H q[0];, Vdg q[1];, Vdg q[2];, Vdg q[3];]

Unwrapping Boxes could introduce arbitrarily complex structures into a Circuit which could possibly invalidate almost all Predicate s, including GateSetPredicate, ConnectivityPredicate, and NoMidMeasurePredicate. It is hence recommended to apply this early in the compilation procedure, prior to any pass that solves for these constraints.

Optimisations#

Having covered the primary goal of compilation and reduced our Circuit s to a form where they can be run, we find that there are additional techniques we can use to obtain more reliable results by reducing the noise and probability of error. Most Circuit optimisations follow the mantra of “fewer expensive resources gives less opportunity for noise to creep in”, whereby if we find an alternative Circuit that is observationally equivalent in a perfect noiseless setting but uses fewer resources (gates, time, ancilla qubits) then it is likely to perform better in a noisy context (though not always guaranteed).

If we have two Circuit s that are observationally equivalent, we know that replacing one for the other in any context also gives something that is observationally equivalent. The simplest optimisations will take an inefficient pattern, find all matches in the given Circuit and replace them by the efficient alternative. A good example from this class of peephole optimisations is the RemoveRedundancies pass, which looks for a number of easy-to-spot redundant gates, such as zero-parameter rotation gates, gate-inverse pairs, adjacent rotation gates in the same basis, and diagonal rotation gates followed by measurements.

from pytket import Circuit, OpType
from pytket.passes import RemoveRedundancies

circ = Circuit(3, 3)
circ.Rx(0.92, 0).CX(1, 2).Rx(-0.18, 0)  # Adjacent Rx gates can be merged
circ.CZ(0, 1).Ry(0.11, 2).CZ(0, 1)      # CZ is self-inverse
circ.XXPhase(0.6, 0, 1)
circ.YYPhase(0, 0, 1)    # 0-angle rotation does nothing
circ.ZZPhase(-0.84, 0, 1)
circ.Rx(0.03, 0).Rz(-0.9, 1).measure_all()  # Effect of Rz is eliminated by measurement

RemoveRedundancies().apply(circ)
print(circ.get_commands())
[Rx(0.74) q[0];, CX q[1], q[2];, XXPhase(0.6) q[0], q[1];, Ry(0.11) q[2];, Measure q[2] --> c[2];, ZZPhase(3.16) q[0], q[1];, Measure q[1] --> c[1];, Rx(0.03) q[0];, Measure q[0] --> c[0];]

It is understandable to question the relevance of such an optimisation, since a sensible programmer would not intentionally write a Circuit with such redundant gates. These are still largely useful because other compiler passes might introduce them, such as routing adding a OpType.SWAP gate immediately following a OpType.SWAP gate made by the user, or commuting a Z-rotation through the control of a CX which allows it to merge with another Z-rotation on the other side.

Previous iterations of the CliffordSimp pass would work in this way as well, looking for specific sequences of Clifford gates where we could reduce the number of two-qubit gates. This has since been generalised to spot these patterns up to gate commutations and changes of basis from single-qubit Clifford rotations.

from pytket import Circuit, OpType
from pytket.passes import CliffordSimp

# A basic inefficient pattern can be reduced by 1 CX
simple_circ = Circuit(2)
simple_circ.CX(0, 1).S(1).CX(1, 0)

CliffordSimp().apply(simple_circ)
print(simple_circ.get_commands())

# The same pattern, up to commutation and local Clifford algebra
complex_circ = Circuit(3)
complex_circ.CX(0, 1)
complex_circ.Rx(0.42, 1)
complex_circ.S(1)
complex_circ.YYPhase(0.96, 1, 2)  # Requires 2 CXs to implement
complex_circ.CX(0, 1)

CliffordSimp().apply(complex_circ)
print(complex_circ.get_commands())
[CX q[1], q[0];, TK1(0, 0, 0.5) q[0];]
[TK1(0, 0, 0.5) q[0];, TK1(0, 0.08, 1) q[1];, TK1(0, 0, 1) q[2];, CX q[2], q[1];, TK1(0.5, 3.04, 0.5) q[2];, CX q[2], q[1];, TK1(0, 0, 0.5) q[1];, CX q[0], q[1];, TK1(0.5, 0.5, 0.5) q[1];]

The next step up in scale has optimisations based on optimal decompositions of subcircuits over \(n\)-qubits, including EulerAngleReduction for single-qubit unitary chains (producing three rotations in a choice of axes), and KAKDecomposition for two-qubit unitaries (using at most three CXs and some single-qubit gates).

from pytket import Circuit, OpType
from pytket.passes import EulerAngleReduction, KAKDecomposition

circ = Circuit(2)
circ.CZ(0, 1)
circ.Rx(0.4, 0).Ry(0.289, 0).Rx(-0.34, 0).Ry(0.12, 0).Rx(-0.81, 0)
circ.CX(1, 0)

# Reduce long chain to a triple of Ry, Rx, Ry
EulerAngleReduction(OpType.Rx, OpType.Ry).apply(circ)
print(circ.get_commands())

circ = Circuit(3)
circ.CX(0, 1)
circ.CX(1, 2).Rx(0.3, 1).CX(1, 2).Rz(1.5, 2).CX(1, 2).Ry(-0.94, 1).Ry(0.37, 2).CX(1, 2)
circ.CX(1, 0)

# Reduce long 2-qubit subcircuit to at most 3 CXs
KAKDecomposition().apply(circ)
print(circ.get_commands())
[CZ q[0], q[1];, Ry(3.88883) q[0];, Rx(3.2547) q[0];, Ry(3.56808) q[0];, CX q[1], q[0];]
[CX q[0], q[1];, TK1(0, 2.5, 3) q[2];, TK1(1.10719, 2.5, 0) q[1];, CX q[1], q[2];, TK1(0, 1.425, 3.5) q[1];, TK1(0, 1.5, 1.94) q[2];, CX q[1], q[2];, TK1(0, 2.5, 3.81281) q[1];, TK1(0, 0, 1.5) q[2];, CX q[1], q[0];]

All of these so far are generic optimisations that work for any application, but only identify local redundancies since they are limited to working up to individual gate commutations. Other techniques instead focus on identifying macroscopic structures in a Circuit or convert it entirely into an alternative algebraic representation, and then using the properties of the structures/algebra to find simplifications and resynthesise into basic gates. For example, the PauliSimp pass will represent the entire Circuit as a sequence of exponentials of Pauli operators, capturing the effects of non-Clifford gates as rotations in a basis determined by the Clifford gates. This abstracts away any redundant information in the Clifford gates entirely, and can be used to merge non-Clifford gates that cannot be brought together from any sequence of commutations, as well as finding efficient Clifford constructions for the basis changes.

from pytket import Circuit
from pytket.passes import PauliSimp
from pytket.utils import Graph

circ = Circuit(3)
circ.Rz(0.2, 0)
circ.Rx(0.35, 1)
circ.V(0).H(1).CX(0, 1).CX(1, 2).Rz(-0.6, 2).CX(1, 2).CX(0, 1).Vdg(0).H(1)
circ.H(1).H(2).CX(0, 1).CX(1, 2).Rz(0.8, 2).CX(1, 2).CX(0, 1).H(1).H(2)
circ.Rx(0.1, 1)

PauliSimp().apply(circ)
Graph(circ).get_DAG()
_images/manual_compiler_15_0.svg

This can give great benefits for Circuit s where non-Clifford gates are sparse and there is hence a lot of redundancy in the Clifford change-of-basis sections. But if the Circuit already has a very efficient usage of Clifford gates, this will be lost when converting to the abstract representation, and so the resynthesis is likely to give less efficient sequences. The large structural changes from abstraction and resynthesis can also make routing harder to perform as the interaction graph of the logical qubits can drastically change. The effectiveness of such optimisations depends on the situation, but can be transformative under the right circumstances.

Some of these optimisation passes have optional parameters to customise the routine slightly. A good example is adapting the PauliSimp pass to have a preference for different forms of OpType.CX decompositions. Setting the cx_config option to CXConfigType.Snake (default) will prefer chains of gates where the target of one becomes the control of the next, whereas CXConfigType.Star prefers using a single qubit as the control for many gates, and CXConfigType.Tree introduces entanglement in a balanced tree form. Each of these has its own benefits and drawbacks that could make it more effective for a particular routine, like CXConfigType.Snake giving circuits that are easier to route on linear nearest-neighbour architectures, CXConfigType.Star allowing any of the gates to commute through to cancel out with others at the start or end of the sequence, and CXConfigType.Tree giving optimal depth on a fully-connected device.

from pytket.circuit import Circuit, PauliExpBox
from pytket.passes import PauliSimp
from pytket.pauli import Pauli
from pytket.transform import CXConfigType
from pytket.utils import Graph

pauli_XYXZYXZZ = PauliExpBox([Pauli.X, Pauli.Y, Pauli.X, Pauli.Z, Pauli.Y, Pauli.X, Pauli.Z, Pauli.Z], 0.42)

circ = Circuit(8)
circ.add_gate(pauli_XYXZYXZZ, [0, 1, 2, 3, 4, 5, 6, 7])

PauliSimp(cx_config=CXConfigType.Snake).apply(circ)
print(circ.get_commands())
Graph(circ).get_qubit_graph()
[H q[0];, V q[1];, H q[2];, V q[4];, H q[5];, CX q[7], q[6];, CX q[6], q[5];, CX q[5], q[4];, CX q[4], q[3];, CX q[3], q[2];, CX q[2], q[1];, CX q[1], q[0];, Rz(0.42) q[0];, CX q[1], q[0];, H q[0];, CX q[2], q[1];, S q[0];, Vdg q[1];, CX q[3], q[2];, H q[0];, S q[1];, H q[2];, CX q[4], q[3];, S q[0];, H q[1];, S q[2];, S q[3];, CX q[5], q[4];, V q[0];, S q[1];, H q[2];, H q[3];, Vdg q[4];, CX q[6], q[5];, V q[1];, S q[2];, S q[3];, S q[4];, H q[5];, CX q[7], q[6];, V q[2];, V q[3];, H q[4];, S q[5];, S q[6];, S q[7];, S q[4];, H q[5];, H q[6];, H q[7];, V q[4];, S q[5];, S q[6];, S q[7];, V q[5];, V q[6];, V q[7];]
_images/manual_compiler_16_1.svg
PauliSimp(cx_config=CXConfigType.Star).apply(circ)
print(circ.get_commands())
Graph(circ).get_qubit_graph()
[H q[0];, V q[1];, H q[2];, V q[4];, H q[5];, CX q[7], q[0];, CX q[6], q[0];, CX q[5], q[0];, CX q[4], q[0];, CX q[3], q[0];, CX q[2], q[0];, CX q[1], q[0];, Rz(0.42) q[0];, CX q[1], q[0];, CX q[2], q[0];, Vdg q[1];, CX q[3], q[0];, S q[1];, H q[2];, CX q[4], q[0];, H q[1];, S q[2];, S q[3];, CX q[5], q[0];, S q[1];, H q[2];, H q[3];, Vdg q[4];, CX q[6], q[0];, V q[1];, S q[2];, S q[3];, S q[4];, H q[5];, CX q[7], q[0];, V q[2];, V q[3];, H q[4];, S q[5];, S q[6];, H q[0];, S q[4];, H q[5];, H q[6];, S q[7];, S q[0];, V q[4];, S q[5];, S q[6];, H q[7];, H q[0];, V q[5];, V q[6];, S q[7];, S q[0];, V q[7];, V q[0];]
_images/manual_compiler_17_1.svg
PauliSimp(cx_config=CXConfigType.Tree).apply(circ)
print(circ.get_commands())
Graph(circ).get_qubit_graph()
[H q[0];, V q[1];, H q[2];, V q[4];, H q[5];, CX q[7], q[6];, CX q[1], q[0];, CX q[3], q[2];, CX q[5], q[4];, CX q[2], q[0];, CX q[6], q[4];, CX q[4], q[0];, Rz(0.42) q[0];, CX q[4], q[0];, CX q[2], q[0];, CX q[6], q[4];, CX q[1], q[0];, CX q[3], q[2];, CX q[5], q[4];, CX q[7], q[6];, H q[0];, Vdg q[1];, H q[2];, S q[3];, Vdg q[4];, H q[5];, S q[6];, S q[7];, S q[0];, S q[1];, S q[2];, H q[3];, S q[4];, S q[5];, H q[6];, H q[7];, H q[0];, H q[1];, H q[2];, S q[3];, H q[4];, H q[5];, S q[6];, S q[7];, S q[0];, S q[1];, S q[2];, V q[3];, S q[4];, S q[5];, V q[6];, V q[7];, V q[0];, V q[1];, V q[2];, V q[4];, V q[5];]
_images/manual_compiler_18_1.svg

Combinators#

The passes encountered so far represent elementary, self-contained transformations on Circuit s. In practice, we will almost always want to apply sequences of these to combine optimisations with solving for many constraints. The passes in pytket have a rudimentary compositional structure to describe generic compilation strategies, with the most basic example being just applying a list of passes in order.

from pytket import Circuit, OpType
from pytket.passes import auto_rebase_pass, EulerAngleReduction, SequencePass

rebase_quil = auto_rebase_pass({OpType.CZ, OpType.Rz, OpType.Rx})
circ = Circuit(3)
circ.CX(0, 1).Rx(0.3, 1).CX(2, 1).Rz(0.8, 1)
comp = SequencePass([rebase_quil, EulerAngleReduction(OpType.Rz, OpType.Rx)])
comp.apply(circ)
print(circ.get_commands())
[Rz(1.5) q[1];, Rx(3.5) q[1];, CZ q[0], q[1];, Rz(2.3) q[1];, CZ q[2], q[1];, Rx(0.5) q[1];, Rz(1.3) q[1];]

When composing optimisation passes, we may find that applying one type of optimisation could open up opportunities for others by, for example, rearranging gates to match the desired template. To make the most of this, it may be beneficial to apply some pass combination repeatedly until no further changes are made, i.e. until we have found and exploited every simplification that we can.

from pytket import Circuit
from pytket.passes import RemoveRedundancies, CommuteThroughMultis, RepeatPass, SequencePass

circ = Circuit(4)
circ.CX(2, 3).CY(1, 2).CX(0, 1).Rz(0.24, 0).CX(0, 1).Rz(0.89, 1).CY(1, 2).Rz(-0.3, 2).CX(2, 3)
comp = RepeatPass(SequencePass([CommuteThroughMultis(), RemoveRedundancies()]))
comp.apply(circ)
print(circ.get_commands())
[Rz(0.24) q[0];, Rz(0.89) q[1];, Rz(3.7) q[2];]

Warning

This looping mechanism does not directly compare the Circuit to its old state from the previous iteration, instead checking if any of the passes within the loop body claimed they performed any rewrite. Some sequences of passes will do and undo some changes to the Circuit, giving no net effect but nonetheless causing the loop to repeat. This can lead to infinite loops if used in such a way. Some passes where the Circuit is converted to another form and back again (e.g. PauliSimp) will always report that a change took place. We recommend testing any looping passes thoroughly to check for termination.

Increased termination safety can be given by only repeating whilst some easy-to-check metric (such as number of gates or depth) decreases. For example, we may want to try to minimise the number of OpType.CX gates since these will tend to be very slow and noisy on a lot of devices.

from pytket import Circuit, OpType
from pytket.passes import RemoveRedundancies, CommuteThroughMultis, RepeatWithMetricPass, SequencePass

circ = Circuit(4)
circ.CX(2, 3).CY(1, 2).CX(0, 1).Rz(0.24, 0).CX(0, 1).Rz(0.89, 1).CY(1, 2).Rz(-0.3, 2).CX(2, 3)
cost = lambda c : c.n_gates_of_type(OpType.CX)
comp = RepeatWithMetricPass(SequencePass([CommuteThroughMultis(), RemoveRedundancies()]), cost)
comp.apply(circ)  # Stops earlier than before, since removing CYs doesn't change the number of CXs
print(circ.get_commands())
[Rz(0.24) q[0];, Rz(0.89) q[1];, CX q[2], q[3];, Rz(3.7) q[2];, CX q[2], q[3];]

We mentioned earlier that each pass has a set of pre-conditions and post-conditions expressed via Predicate s. We may find that applying one pass invalidates the pre-conditions of a later pass, meaning it may hit an error when applied to a Circuit. For example, the KAKDecomposition optimisation method can only operate on Circuit s with a specific gate set which doesn’t allow for any gates on more than 2 qubits, so when RoutingPass can introduce OpType.BRIDGE gates over 3 qubits, this could cause an error when trying to apply KAKDecomposition. When using combinators like SequencePass and RepeatPass, pytket checks that the passes are safe to compose, in the sense that former passes do not invalidate pre-conditions of the latter passes. This procedure uses a basic form of Hoare logic to identify new pre- and post-conditions for the combined pass and identify whether it is still satisfiable.

A special mention here goes to the DecomposeBoxes pass. Because the Box structures could potentially contain arbitrary sequences of gates, there is no guarantee that expanding them will yield a Circuit that satisfies any Predicate. Since it has potential to invalidate the pre-conditions of any subsequent pass, composing it with anything else will generate such an error.

from pytket.passes import DecomposeBoxes, PauliSimp, SequencePass
# PauliSimp requires a specific gateset and no conditional gates
# or mid-circuit measurement, so this will raise an exception
comp = SequencePass([DecomposeBoxes(), PauliSimp()])
---------------------------------------------------------------------------
RuntimeError                              Traceback (most recent call last)
Cell In[19], line 4
      1 from pytket.passes import DecomposeBoxes, PauliSimp, SequencePass
      2 # PauliSimp requires a specific gateset and no conditional gates
      3 # or mid-circuit measurement, so this will raise an exception
----> 4 comp = SequencePass([DecomposeBoxes(), PauliSimp()])

RuntimeError: Cannot compose these Compiler Passes due to mismatching Predicates of type: GateSetPredicate

Predefined Sequences#

Knowing what sequences of compiler passes to apply for maximal performance is often a very hard problem and can require a lot of experimentation and intuition to predict reliably. Fortunately, there are often common patterns that are applicable to virtually any scenario, for which pytket provides some predefined sequences.

In practice, peephole and structure-preserving optimisations are almost always strictly beneficial to apply, or at least will never increase the size of the Circuit. The FullPeepholeOptimise pass applies Clifford simplifications, commutes single-qubit gates to the front of the circuit and applies passes to squash subcircuits of up to three qubits. This provides a one-size-approximately-fits-all “kitchen sink” solution to Circuit optimisation. This assumes no device constraints by default, so will not generally preserve gateset, connectivity, etc.

When targeting a heterogeneous device architecture, solving this constraint in its entirety will generally require both placement and subsequent routing. DefaultMappingPass simply combines these to apply the GraphPlacement strategy and solve any remaining invalid multi-qubit operations. This is taken a step further with CXMappingPass which also decomposes the introduced OpType.SWAP and OpType.BRIDGE gates into elementary OpType.CX gates.

After solving for the device connectivity, we then need to restrict what optimisations we can apply to those that won’t invalidate this. The set of SynthesiseX passes combine light optimisations that preserve the qubit connectivity and target a specific final gate set (e.g. SynthesiseTket guarantees the output is in the gateset of OpType.CX, OpType.TK1, and OpType.Measure). In general, this will not reduce the size of a Circuit as much as FullPeepholeOptimise, but has the benefit of removing some redundancies introduced by routing without invalidating it.

from pytket import Circuit, OpType
from pytket.extensions.qiskit import IBMQBackend
from pytket.passes import FullPeepholeOptimise, DefaultMappingPass, SynthesiseTket, RebaseTket

circ = Circuit(5)
circ.CX(0, 1).CX(0, 2).CX(0, 3)
circ.CZ(0, 1).CZ(0, 2).CZ(0, 3)
circ.CX(3, 4).CX(0, 3).CX(4, 0)

RebaseTket().apply(circ)     # Get number of 2qb gates by converting all to CX
print(circ.n_gates_of_type(OpType.CX))

FullPeepholeOptimise().apply(circ)      # Freely rewrite circuit
print(circ.n_gates_of_type(OpType.CX))

backend = IBMQBackend("ibmq_quito")
DefaultMappingPass(backend.backend_info.architecture).apply(circ)
RebaseTket().apply(circ)
print(circ.n_gates_of_type(OpType.CX))  # Routing adds gates
print(circ.get_commands())

SynthesiseTket().apply(circ)            # Some added gates may be redundant
print(circ.n_gates_of_type(OpType.CX))  # But not in this case
9
6
9
[tk1(0, 0, 1.5) node[0];, tk1(0, 0, 1.5) node[1];, tk1(0, 0, 1.5) node[2];, tk1(0, 0, 1.5) node[3];, CX node[1], node[0];, tk1(0, 0, 0.5) node[0];, CX node[1], node[2];, CX node[1], node[3];, tk1(0, 0, 0.5) node[2];, tk1(0, 0, 0.5) node[3];, CX node[3], node[4];, CX node[1], node[3];, CX node[3], node[4];, CX node[4], node[3];, CX node[3], node[4];, CX node[3], node[1];]
9

Note

FullPeepholeOptimise takes an optional allow_swaps argument. This is a boolean flag to indicate whether FullPeepholeOptimise should preserve the circuit connectivity or not. If set to False the pass will presrve circuit connectivity but the circuit will generally be less optimised than if connectivity was ignored.

FullPeepholeOptimise also takes an optional target_2qb_gate argument to specify whether to target the {OpType.TK1, OpType.CX} or {OpType.TK1, OpType.TK2} gateset.

Note

Prevous versions of FullPeepholeOptimise did not apply the ThreeQubitSquash pass. There is a PeepholeOptimise2Q pass which applies the old pass sequence with the ThreeQubitSquash pass excluded.

Also in this category of pre-defined sequences, we have the default_compilation_pass() which is run by get_compiled_circuit(). These give a recommended compiler pass to solve the Backend s constraints with a choice of optimisation levels.

Optimisation level

Description

0

Just solves the constraints as simply as possible. No optimisation.

1

Adds basic optimisations (those covered by the SynthesiseX passes) for efficient compilation.

2

Extends to more intensive optimisations (those covered by the FullPeepholeOptimise pass).

We will now demonstrate the default_compilation_pass() with the different levels of optimisation using the ibmq_quito device.

As more intensive optimisations are applied by level 2 the pass may take a long to run for large circuits. In this case it may be preferable to apply the lighter optimisations of level 1.

from pytket import Circuit, OpType
from pytket.extensions.qiskit import IBMQBackend

circ = Circuit(3) # Define a circuit to be compiled to the backend
circ.CX(0, 1)
circ.H(1)
circ.Rx(0.42, 1)
circ.S(1)
circ.CX(0, 2)
circ.CX(2, 1)
circ.Z(2)
circ.Y(1)
circ.CX(0, 1)
circ.CX(2, 0)
circ.measure_all()

backend = IBMQBackend("ibmq_quito") # Initialise Backend

print("Total gate count before compilation =", circ.n_gates)
print("CX count before compilation =",  circ.n_gates_of_type(OpType.CX))

 # Now apply the default_compilation_pass at different levels of optimisation.

for ol in range(3):
    test_circ = circ.copy()
    backend.default_compilation_pass(optimisation_level=ol).apply(test_circ)
    assert backend.valid_circuit(test_circ)
    print("Optimisation level", ol)
    print("Gates", test_circ.n_gates)
    print("CXs", test_circ.n_gates_of_type(OpType.CX))
Total gate count before compilation = 13
CX count before compilation = 5
Optimisation level 0
Gates 22
CXs 8
Optimisation level 1
Gates 12
CXs 5
Optimisation level 2
Gates 6
CXs 1

Explanation

We see that compiling the circuit to ibmq_quito at optimisation level 0 actually increases the gate count. This is because ibmq_quito has connectivity constraints which require additional CX gates to be added to validate the circuit. The single-qubit gates in our circuit also need to be decomposed into the IBM gatset.

We see that compiling at optimisation level 1 manages to reduce the CX count to 5. Our connectivity constraints are satisfied without increasing the CX gate count. Single-qubit gates are also combined to reduce the overall gate count further.

Finally we see that the default pass for optimisation level 2 manages to reduce the overall gate count to just 6 with only one CX gate. This is because more intensive optimisations are applied at this level including squashing passes that enable optimal two and three-qubit circuits to be synthesised. Applying these more powerful passes comes with a runtime overhead that may be noticeable for larger circuits.

Guidance for Combining Passes#

We find that the most powerful optimisation techniques (those that have the potential to reduce Circuit size the most for some class of Circuit s) tend to have fewer guarantees on the structure of the output, requiring a universal quantum computer with the ability to perform any gates on any qubits. It is recommended to apply these early on in compilation.

The passes to solve some device constraints might invalidate others: for example, the RoutingPass generally invalidates NoMidMeasurePredicate and GateSetPredicate. Therefore, the order in which these are solved should be chosen with care.

For most standard use cases, we recommend starting with DecomposeBoxes to reduce the Circuit down to primitive gates, followed by strong optimisation passes like PauliSimp (when appropriate for the types of Circuit s being considered) and FullPeepholeOptimise to eliminate a large number of redundant operations. Then start to solve some more device constraints with some choice of placement and routing strategy, followed by DelayMeasures to push measurements back through any introduced OpType.SWAP or OpType.BRIDGE gates, and then finally rebase to the desired gate set. The default_compilation_pass() definitions can replace this sequence from placement onwards for simplicity. Minor optimisations could also be inserted between successive steps to tidy up any redundancies introduced, as long as they preserve the solved constraints.

Initial and Final Maps#

PlacementPass modifies the set of qubits used in the Circuit from the logical names used during construction to the names of the physical addresses on the Backend, so the logical qubit names wiil no longer exist within the Circuit by design. Knowing the map between the logical qubits and the chosen physical qubits is necessary for understanding the choice of placement, interpreting the final state from a naive simulator, identifying which physical qubits each measurement was made on for error mitigation, and appending additional gates to the logical qubits after applying the pass.

Other passes like RoutingPass and CliffordSimp can introduce (explicit or implicit) permutations of the logical qubits in the middle of a Circuit, meaning a logical qubit may exist on a different physical qubit at the start of the Circuit compared to the end.

We can wrap up a Circuit in a CompilationUnit to allow us to track any changes to the locations of the logical qubits when passes are applied. The CompilationUnit.initial_map is a dictionary mapping the original UnitID s to the corresponding UnitID used in CompilationUnit.circuit, and similarly CompilationUnit.final_map for outputs. Applying BasePass.apply() to a CompilationUnit will apply the transformation to the underlying Circuit and track the changes to the initial and final maps.

from pytket import Circuit
from pytket.extensions.qiskit import IBMQBackend
from pytket.passes import DefaultMappingPass
from pytket.predicates import CompilationUnit

circ = Circuit(5, 5)
circ.CX(0, 1).CX(0, 2).CX(0, 3).CX(0, 4).measure_all()
backend = IBMQBackend("ibmq_quito")
cu = CompilationUnit(circ)
DefaultMappingPass(backend.backend_info.architecture).apply(cu)
print(cu.circuit.get_commands())
print(cu.initial_map)
print(cu.final_map)
[CX node[1], node[0];, Measure node[0] --> c[1];, CX node[1], node[2];, Measure node[2] --> c[2];, CX node[1], node[3];, Measure node[3] --> c[3];, SWAP node[1], node[3];, CX node[3], node[4];, Measure node[3] --> c[0];, Measure node[4] --> c[4];]
{c[0]: c[0], c[1]: c[1], c[2]: c[2], c[3]: c[3], c[4]: c[4], q[0]: node[1], q[1]: node[0], q[2]: node[2], q[3]: node[3], q[4]: node[4]}
{c[0]: c[0], c[1]: c[1], c[2]: c[2], c[3]: c[3], c[4]: c[4], q[0]: node[3], q[1]: node[0], q[2]: node[2], q[3]: node[1], q[4]: node[4]}

Note

No passes currently rename or swap classical data, but the classical bits are included in these maps for completeness.

Advanced Compilation Topics#

Compiling Symbolic Circuits#

For variational algorithms, the prominent benefit of defining a Circuit symbolically and only instantiating it with concrete values when needed is that the compilation procedure would only need to be performed once. By saving time here we can cut down the overall time for an experiment; we could invest the time saved into applying more expensive optimisations on the Circuit to reduce the impact of noise further.

from pytket import Circuit, Qubit
from pytket.extensions.qiskit import AerStateBackend
from pytket.pauli import Pauli, QubitPauliString
from pytket.utils.operators import QubitPauliOperator
from sympy import symbols

a, b = symbols("a b")
circ = Circuit(2)
circ.Ry(a, 0)
circ.Ry(a, 1)
circ.CX(0, 1)
circ.Rz(b, 1)
circ.CX(0, 1)
xx = QubitPauliString({Qubit(0):Pauli.X, Qubit(1):Pauli.X})
op = QubitPauliOperator({xx : 1.5})

backend = AerStateBackend()

# Compile once outside of the objective function
circ = backend.get_compiled_circuit(circ)

def objective(params):
    state = circ.copy()
    state.symbol_substitution({a : params[0], b : params[1]})
    handle = backend.process_circuit(state) # No need to compile again
    vec = backend.get_result(handle).get_state()
    return op.state_expectation(vec)

print(objective([0.25, 0.5]))
print(objective([0.5, 0]))
(0.75+0j)
(1.5+0j)

Note

Every Backend requires NoSymbolsPredicate, so it is necessary to instantiate all symbols before running a Circuit.

User-defined Passes#

We have already seen that pytket allows users to combine passes in a desired order using SequencePass. An addtional feature is the CustomPass which allows users to define their own custom circuit transformation using pytket. The CustomPass class accepts a transform parameter, a python function that takes a Circuit as input and returns a Circuit as output.

We will show how to use CustomPass by defining a simple transformation that replaces any Pauli Z gate in the Circuit with a Hadamard gate, Pauli X gate, Hadamard gate chain.

from pytket import Circuit, OpType

def z_transform(circ: Circuit) -> Circuit:
    n_qubits = circ.n_qubits
    circ_prime = Circuit(n_qubits) # Define a replacement circuit

    for cmd in circ.get_commands():
        qubit_list = cmd.qubits # Qubit(s) our gate is applied on (as a list)
        if cmd.op.type == OpType.Z:
            # If cmd is a Z gate, decompose to a H, X, H sequence.
            circ_prime.add_gate(OpType.H, qubit_list)
            circ_prime.add_gate(OpType.X, qubit_list)
            circ_prime.add_gate(OpType.H, qubit_list)
        else:
            # Otherwise, apply the gate as usual.
            circ_prime.add_gate(cmd.op.type, cmd.op.params, qubit_list)

    return circ_prime

After we’ve defined our transform we can construct a CustomPass. This pass can then be applied to a Circuit.

from pytket.passes import CustomPass

DecompseZPass = CustomPass(z_transform) # Define our pass

test_circ = Circuit(2) # Define a test Circuit for our pass
test_circ.Z(0)
test_circ.Z(1)
test_circ.CX(0, 1)
test_circ.Z(1)
test_circ.CRy(0.5, 0, 1)

DecompseZPass.apply(test_circ) # Apply our pass to the test Circuit

test_circ.get_commands() # Commands of our transformed Circuit
[H q[0];,
 H q[1];,
 X q[0];,
 X q[1];,
 H q[0];,
 H q[1];,
 CX q[0], q[1];,
 H q[1];,
 X q[1];,
 H q[1];,
 CRy(0.5) q[0], q[1];]

We see from the output above that our newly defined DecompseZPass has successfully decomposed the Pauli Z gates to Hadamard, Pauli X, Hadamard chains and left other gates unchanged.

Warning

pytket does not require that CustomPass preserves the unitary of the Circuit . This is for the user to ensure.

Partial Compilation#

A common pattern across expectation value and tomography experiments is to run many Circuit s that have large identical regions, such as a single state preparation with many different measurements. We can further speed up the overall compilation time by splitting up the state preparation from the measurements, compiling each subcircuit only once, and composing together at the end.

The main technical consideration here is that the compiler will only have the freedom to identify good placements for the first subcircuit to be run. This means that the state preparation should be compiled first, and the placement for the measurements is given by the final map in order to compose well.

Once compiled, we can use process_circuits() to submit several circuits at once for execution on the backend. The circuits to be executed are passed as list. If the backend is shot-based, the number of shots can be passed using the n_shots parameter, which can be a single integer or a list of integers of the same length as the list of circuits to be executed. In the following example, 4000 shots are measured for the first circuit and 2000 for the second.

from pytket import Circuit, OpType
from pytket.extensions.qiskit import IBMQBackend
from pytket.predicates import CompilationUnit
from pytket.placement import Placement

state_prep = Circuit(4)
state_prep.H(0)
state_prep.add_gate(OpType.CnRy, 0.1, [0, 1])
state_prep.add_gate(OpType.CnRy, 0.2, [0, 2])
state_prep.add_gate(OpType.CnRy, 0.3, [0, 3])

measure0 = Circuit(4, 4)
measure0.H(1).H(3).measure_all()
measure1 = Circuit(4, 4)
measure1.CX(1, 2).CX(3, 2).measure_all()

backend = IBMQBackend("ibmq_quito")
cu = CompilationUnit(state_prep)
backend.default_compilation_pass().apply(cu)
Placement.place_with_map(measure0, cu.final_map)
Placement.place_with_map(measure1, cu.final_map)
backend.default_compilation_pass().apply(measure0)
backend.default_compilation_pass().apply(measure1)

circ0 = cu.circuit
circ1 = circ0.copy()
circ0.append(measure0)
circ1.append(measure1)
handles = backend.process_circuits([circ0, circ1], n_shots=[4000, 2000])
r0, r1 = backend.get_results(handles)
print(r0.get_counts())
print(r1.get_counts())
{(0, 0, 0, 0): 503, (0, 0, 0, 1): 488, (0, 1, 0, 0): 533, (0, 1, 0, 1): 493, (1, 0, 0, 0): 1041, (1, 0, 0, 1): 107, (1, 0, 1, 0): 115, (1, 0, 1, 1): 14, (1, 1, 0, 0): 576, (1, 1, 0, 1): 69, (1, 1, 1, 0): 54, (1, 1, 1, 1): 7}
{(0, 0, 0, 0): 2047, (0, 1, 0, 0): 169, (0, 1, 1, 0): 1729, (1, 1, 0, 0): 7, (1, 1, 1, 0): 48}

Measurement Reduction#

Suppose we have one of these measurement scenarios (i.e. a single state preparation, but many measurements to make on it) and that each of the measurements is a Pauli observable, such as when calculating the expectation value of the state with respect to some QubitPauliOperator. Naively, we would need a different measurement Circuit per term in the operator, but we can reduce this by exploiting the fact that commuting observables can be measured simultaneously.

Given a set of observables, we can partition them into subsets that are easy to measure simultaneously. A Circuit is generated for each subset by diagonalising the observables (reducing all of them to a combination of \(Z\)-measurements).

Diagonalising a mutually commuting set of Pauli observables could require an arbitrary Clifford circuit in general. If we are considering the near-term regime where “every gate counts”, the diagonalisation of the observables could introduce more of the (relatively) expensive two-qubit gates, giving us the speedup at the cost of some extra noise. pytket can partition the Pauli observables into either general commuting sets for improved reduction in the number of measurement Circuit s, or into smaller sets which can be diagonalised without introducing any multi-qubit gates - this is possible when all observables are substrings of some measured Pauli string (e.g. XYI and IYZ is fine, but ZZZ and XZX is not).

This measurement partitioning is built into the get_operator_expectation_value() utility method, or can be used directly using pytket.partition.measurement_reduction() which builds a MeasurementSetup object. A MeasurementSetup contains a list of measurement Circuit s and a map from the QubitPauliString of each observable to the information required to extract the expectation value (which bits to consider from which Circuit).

from pytket import Qubit
from pytket.pauli import Pauli, QubitPauliString
from pytket.partition import measurement_reduction, PauliPartitionStrat

zi = QubitPauliString({Qubit(0):Pauli.Z})
iz = QubitPauliString({Qubit(1):Pauli.Z})
zz = QubitPauliString({Qubit(0):Pauli.Z, Qubit(1):Pauli.Z})
xx = QubitPauliString({Qubit(0):Pauli.X, Qubit(1):Pauli.X})
yy = QubitPauliString({Qubit(0):Pauli.Y, Qubit(1):Pauli.Y})

setup = measurement_reduction([zi, iz, zz, xx, yy], strat=PauliPartitionStrat.CommutingSets)
print("Via Commuting Sets:")
for i, c in enumerate(setup.measurement_circs):
    print(i, c.get_commands())
print(setup.results[yy])

setup = measurement_reduction([zi, iz, zz, xx, yy], strat=PauliPartitionStrat.NonConflictingSets)
print("Via Non-Conflicting Sets:")
for i, c in enumerate(setup.measurement_circs):
    print(i, c.get_commands())
print(setup.results[yy])
Via Commuting Sets:
0 [Measure q[0] --> c[0];, Measure q[1] --> c[1];]
1 [CX q[0], q[1];, Measure q[1] --> c[1];, H q[0];, Measure q[0] --> c[0];]
[Circuit index: 1
Bits: 0 1 
Invert: True]
Via Non-Conflicting Sets:
0 [Measure q[0] --> c[0];, Measure q[1] --> c[1];]
1 [H q[0];, H q[1];, Measure q[0] --> c[0];, Measure q[1] --> c[1];]
2 [V q[0];, V q[1];, Measure q[0] --> c[0];, Measure q[1] --> c[1];]
[Circuit index: 2
Bits: 0 1 
Invert: False]

Note

Since there could be multiple measurement Circuit s generating the same observable, we could theoretically use this to extract extra shots (and hence extra precision) for that observable for free; automatically doing this as part of measurement_reduction() is planned for a future release of pytket.

Contextual Optimisations#

By default, tket makes no assumptions about a circuit’s input state, nor about the destiny of its output state. We can therefore compose circuits freely, construct boxes from them that we can then place inside other circuits, and so on. However, when we come to run a circuit on a real device we can almost always assume that it will be initialised in the all-zero state, and that the final state of the qubits will be discarded (after measurement).

This is where contextual optimisations can come into play. These are optimisations that depend on knowledge of the context of the circuit being run. They do not generally preserve the full unitary, but they generate circuits that are observationally indistinguishable (on an ideal device), and reduce noise by eliminating unnecessary operations from the beginning or end of the circuit.

First of all, tket provides methods to annotate a qubit (or all qubits) as being initialized to zero, or discarded at the end of the circuit, or both.

from pytket import Circuit

c = Circuit(2)
c.Y(0)
c.CX(0,1)
c.H(0)
c.H(1)
c.Rz(0.125, 1)
c.measure_all()
c.qubit_create_all()
c.qubit_discard_all()

The last two lines tell the compiler that all qubits are to be initialized to zero and discarded at the end. The methods Circuit.qubit_create() and Circuit.qubit_discard() can be used to achieve the same on individual qubits.

Warning

Note that we are now restricted in how we can compose our circuit with other circuits. When composing after another circuit, a “created” qubit becomes a Reset operation. Whem composing before another circuit, a “discarded” qubit may not be joined to another qubit unless that qubit has itself been “created” (so that the discarded state gets reset to zero).

Initial simplification#

When the above circuit is run from an all-zero state, the Y and CX gates at the beginning just have the effect of putting both qubits in the \(\lvert 1 \rangle\) state (ignoring unobservable global phase), so they could be replaced with two X gates. This is exactly what the SimplifyInitial pass does.

from pytket.passes import SimplifyInitial

SimplifyInitial().apply(c)
print(c.get_commands())
[X q[0];, X q[1];, H q[0];, H q[1];, Measure q[0] --> c[0];, Measure q[1] --> c[1];]

This pass tracks the state of qubits known to be initialised to zero (or reset mid-circuit) forward through the circuit, for as long as the qubits remain in a computational basis state, either removing gates (when they don’t change the state) or replacing them with X gates (when they invert the state).

By default, this pass also replaces Measure operations acting on qubits with a known state by classical set-bits operations on the target bits:

c = Circuit(1).X(0).measure_all()
c.qubit_create_all()
SimplifyInitial().apply(c)
print(c.get_commands())
[SetBits(1) c[0];, X q[0];]

The measurement has disappeared, replaced with a classical operation on its target bit. To disable this behaviour, pass the allow_classical=False argument to SimplifyInitial when constructing the pass.

Warning

Most backends currently do not support set-bit operations, so these could cause errors when using this pass with mid-circuit measurements. In such cases you should set allow_classical=False.

Note that SimplifyInitial does not automatically cancel successive pairs of X gates introduced by the simplification. It is a good idea to follow it with a RemoveRedundancies pass in order to perform these cancellations.

Removal of discarded operations#

An operation that has no quantum or classical output in its causal future has no effect (or rather, no observable effect on an ideal system), and can be removed from the circuit. By marking a qubit as discarded, we tell the compiler that it has no quantum output, potentially enabling this simplification.

Note that if the qubit is measured, even if it is then discarded, the Measure operation has a classical output in its causal future so will not be removed.

from pytket.circuit import Qubit
from pytket.passes import RemoveDiscarded

c = Circuit(3, 2)
c.H(0).H(1).H(2).CX(0, 1).Measure(0, 0).Measure(1, 1).H(0).H(1)
c.qubit_discard(Qubit(0))
c.qubit_discard(Qubit(2))
RemoveDiscarded().apply(c)
print(c.get_commands())
[H q[0];, H q[1];, CX q[0], q[1];, Measure q[0] --> c[0];, Measure q[1] --> c[1];, H q[1];]

The Hadamard gate following the measurement on qubit 0, as well as the Hadamard on qubit 2, have disappeared, because those qubits were discarded. The Hadamard following the measurement on qubit 1 remains, because that qubit was not discarded.

Commutation of measured classical maps#

The last type of contextual optimization is a little more subtle. Let’s call a quantum unitary operation a classical map if it sends every computational basis state to a computational basis state, possibly composed with a diagonal operator. For example, X, Y, Z, Rz, CX, CY, CZ and Sycamore are classical maps, but Rx, Ry and H are not. Check the documentation of gate types to see which gates have unitaries that make them amenable to optimisation.

When a classical map is followed by a measurement of all its qubits, and those qubits are then discarded, it can be replaced by a purely classical operation acting on the classical outputs of the measurement.

For example, if we apply a CX gate and then measure the two qubits, the result is (ideally) the same as if we measured the two qubits first and then applied a classical controlled-NOT on the measurement bits. If the gate were a CY instead of a CX the effect would be identical: the only difference is the insertion of a diagonal operator, whose effect is unmeasurable.

This simplification is effected by the SimplifyMeasured pass.

Let’s illustrate this with a Bell circuit:

from pytket.passes import SimplifyMeasured

c = Circuit(2).H(0).CX(0, 1).measure_all()
c.qubit_discard_all()
SimplifyMeasured().apply(c)
print(c.get_commands())
[Measure q[1] --> c[1];, H q[0];, Measure q[0] --> c[0];, ClassicalTransform c[0], c[1];]

The CX gate has disappeared, replaced with a classical transform acting on the bits after the measurement.

Contextual optimisation in practice#

The above three passes are combined in the ContextSimp pass, which also performs a final RemoveRedundancies. Normally, before running a circuit on a device you will want to apply this pass (after using Circuit.qubit_create_all() and Circuit.qubit_discard_all() to enable the simplifications).

However, most backends cannot process the classical operations that may be introduced by SimplifyMeasured or (possibly) SimplifyInitial. So pytket provides a method separate_classical() to separate the classical postprocessing circuit from the main circuit to be run on the device. This postprocessing circuit is then passed as the ppcirc argument to get_counts() or get_shots(), in order to obtain the postprocessed results.

Much of the above is wrapped up in the utility method prepare_circuit(). This takes a circuit, applies Circuit.qubit_create_all() and Circuit.qubit_discard_all(), runs the full ContextSimp pass, and then separates the result into the main circuit and the postprocessing circuit, returning both.

Thus a typical usage would look something like this:

from pytket.utils import prepare_circuit
from pytket.extensions.qiskit import AerBackend

b = AerBackend()
c = Circuit(2).H(0).CX(0, 1)
c.measure_all()
c0, ppcirc = prepare_circuit(c)
c0 = b.get_compiled_circuit(c0)
h = b.process_circuit(c0, n_shots=10)
r = b.get_result(h)
shots = r.get_shots(ppcirc=ppcirc)
print(shots)
[[0 0]
 [0 0]
 [0 0]
 [0 0]
 [0 0]
 [0 0]
 [1 1]
 [1 1]
 [1 1]
 [1 1]]

This is a toy example, but illustrates the principle. The actual circuit sent to the backend consisted only of a Hadamard gate on qubit 0 and a single measurement to bit 0. The classical postprocessing circuit set bit 1 to zero and then executed a controlled-NOT from bit 0 to bit 1. These details are hidden from us (unless we inspect the circuits), and what we end up with is a shots table that is indistinguishable from running the original circuit but with less noise.