Preamble

This document aims to describe the purpose, functionality, and implementation of the host for Polkadot's parachains functionality - that is, the software which provides security and advancement for constituent parachains. It is not for the implementer of a specific parachain but rather for the implementer of the Parachain Host. In practice, this is for the implementers of Polkadot in general.

There are a number of other documents describing the research in more detail. All referenced documents will be linked here and should be read alongside this document for the best understanding of the full picture. However, this is the only document which aims to describe key aspects of Polkadot's particular instantiation of much of that research down to low-level technical details and software architecture.

Whence Parachains

Parachains are the solution to a problem. As with any solution, it cannot be understood without first understanding the problem. So let's start by going over the issues faced by blockchain technology that led to us beginning to explore the design space for something like parachains.

Issue 1: Scalability

It became clear a few years ago that the transaction throughput of simple Proof-of-Work (PoW) blockchains such as Bitcoin, Ethereum, and myriad others was simply too low.

TODO: what if there were more blockchains, etc.

Proof-of-Stake (PoS) systems can accomplish higher throughput than PoW blockchains. PoS systems are secured by bonded capital as opposed to spent effort - liquidity opportunity cost vs. burning electricity. The way they work is by selecting a set of validators with known economic identity who lock up tokens in exchange for earning the right to "validate" or participate in the consensus process. If they are found to carry out that process wrongly, they will be slashed, meaning some or all of the locked tokens will be burned. This provides a strong disincentive in the direction of misbehavior.

Since the consensus protocol doesn't revolve around wasting effort, block times and agreement can occur much faster. Solutions to PoW challenges don't have to be found before a block can be authored, so the overhead of authoring a block is reduced to only the costs of creating and distributing the block.

However, consensus on a PoS chain requires full agreement of 2/3+ of the validator set for everything that occurs at Layer 1: all logic which is carried out as part of the blockchain's state machine. This means that everybody still needs to check everything. Furthermore, validators may have different views of the system based on the information that they receive over an asynchronous network, making agreement on the latest state more difficult.

Parachains are an example of a sharded protocol. Sharding is a concept borrowed from traditional database architecture. Rather than requiring every participant to check every transaction, we require each participant to check some subset of transactions, with enough redundancy baked in that byzantine (arbitrarily malicious) participants can't sneak in invalid transactions - at least not without being detected and getting slashed, with those transactions reverted.

Sharding and Proof-of-Stake in coordination with each other allow a parachain host to provide full security on many parachains, even without all participants checking all state transitions.

TODO: note about network effects & bridging

Issue 2: Flexibility / Specialization

"dumb" VMs don't give you the flexibility. Any engineer knows that being able to specialize on a problem gives them and their users more leverage.

TODO: expand on leverage

Having recognized these issues, we set out to find a solution to these problems, which could allow developers to create and deploy purpose-built blockchains unified under a common source of security, with the capability of message-passing between them; a heterogeneous sharding solution, which we have come to know as Parachains.

Protocol Overview

This section aims to describe, at a high level, the actors and protocols involved in running parachains in Polkadot. Specifically, we describe how different actors communicate with each other, what data structures they keep both individually and collectively, and the high-level purpose on why they do these things.

Our top-level goal is to carry a parachain block from authoring to secure inclusion, and define a process which can be carried out repeatedly and in parallel for many different parachains to extend them over time. Understanding of the high-level approach taken here is important to provide context for the proposed architecture further on. The key parts of Polkadot relevant to this are the main Polkadot blockchain, known as the relay-chain, and the actors which provide security and inputs to this blockchain.

First, it's important to go over the main actors we have involved in this protocol.

  1. Validators. These nodes are responsible for validating proposed parachain blocks. They do so by checking a Proof-of-Validity (PoV) of the block and ensuring that the PoV remains available. They put financial capital down as "skin in the game" which can be slashed (destroyed) if they are proven to have misvalidated.
  2. Collators. These nodes are responsible for creating the Proofs-of-Validity that validators know how to check. Creating a PoV typically requires familiarity with the transaction format and block authoring rules of the parachain, as well as having access to the full state of the parachain.

This implies a simple pipeline where collators send validators parachain blocks and their requisite PoV to check. Then, validators validate the block using the PoV, signing statements which describe either the positive or negative outcome, and with enough positive statements, the block can be noted on the relay-chain. Negative statements are not a veto but will lead to a dispute, with those on the wrong side being slashed. If another validator later detects that a validator or group of validators incorrectly signed a statement claiming a block was valid, then those validators will be slashed, with the checker receiving a bounty.

However, there is a problem with this formulation. In order for another validator to check the previous group of validators' work after the fact, the PoV must remain available so the other validator can fetch it in order to check the work. The PoVs are expected to be too large to include in the blockchain directly, so we require an alternate data availability scheme which requires validators to prove that the inputs to their work will remain available, and so their work can be checked. Empirical tests tell us that many PoVs may be between 1 and 10MB during periods of heavy load.

Here is a description of the Inclusion Pipeline: the path a parachain block (or parablock, for short) takes from creation to inclusion:

  1. Validators are selected and assigned to parachains by the Validator Assignment routine.
  2. A collator produces the parachain block, which is known as a parachain candidate or candidate, along with a PoV for the candidate.
  3. The collator forwards the candidate and PoV to validators assigned to the same parachain via the Collator Protocol.
  4. The validators assigned to a parachain at a given point in time participate in the Candidate Backing subsystem to validate candidates that were put forward for validation. Candidates which gather enough signed validity statements from validators are considered "backable". Their backing is the set of signed validity statements.
  5. A relay-chain block author, selected by BABE, can note up to one (1) backable candidate for each parachain to include in the relay-chain block alongside its backing. A backable candidate once included in the relay-chain is considered backed in that fork of the relay-chain.
  6. Once backed in the relay-chain, the parachain candidate is considered to be "pending availability". It is not considered to be included as part of the parachain until it is proven available.
  7. In the following relay-chain blocks, validators will participate in the Availability Distribution subsystem to ensure availability of the candidate. Information regarding the availability of the candidate will be noted in the subsequent relay-chain blocks.
  8. Once the relay-chain state machine has enough information to consider the candidate's PoV as being available, the candidate is considered to be part of the parachain and is graduated to being a full parachain block, or parablock for short.

Note that the candidate can fail to be included in any of the following ways:

  • The collator is not able to propagate the candidate to any validators assigned to the parachain.
  • The candidate is not backed by validators participating in the Candidate Backing Subsystem.
  • The candidate is not selected by a relay-chain block author to be included in the relay chain
  • The candidate's PoV is not considered as available within a timeout and is discarded from the relay chain.

This process can be divided further down. Steps 2 & 3 relate to the work of the collator in collating and distributing the candidate to validators via the Collation Distribution Subsystem. Steps 3 & 4 relate to the work of the validators in the Candidate Backing Subsystem and the block author (itself a validator) to include the block into the relay chain. Steps 6, 7, and 8 correspond to the logic of the relay-chain state-machine (otherwise known as the Runtime) used to fully incorporate the block into the chain. Step 7 requires further work on the validators' parts to participate in the Availability Distribution Subsystem and include that information into the relay chain for step 8 to be fully realized.

This brings us to the second part of the process. Once a parablock is considered available and part of the parachain, it is still "pending approval". At this stage in the pipeline, the parablock has been backed by a majority of validators in the group assigned to that parachain, and its data has been guaranteed available by the set of validators as a whole. Once it's considered available, the host will even begin to accept children of that block. At this point, we can consider the parablock as having been tentatively included in the parachain, although more confirmations are desired. However, the validators in the parachain-group (known as the "Parachain Validators" for that parachain) are sampled from a validator set which contains some proportion of byzantine, or arbitrarily malicious members. This implies that the Parachain Validators for some parachain may be majority-dishonest, which means that (secondary) approval checks must be done on the block before it can be considered approved. This is necessary only because the Parachain Validators for a given parachain are sampled from an overall validator set which is assumed to be up to <1/3 dishonest - meaning that there is a chance to randomly sample Parachain Validators for a parachain that are majority or fully dishonest and can back a candidate wrongly. The Approval Process allows us to detect such misbehavior after-the-fact without allocating more Parachain Validators and reducing the throughput of the system. A parablock's failure to pass the approval process will invalidate the block as well as all of its descendants. However, only the validators who backed the block in question will be slashed, not the validators who backed the descendants.

The Approval Process, at a glance, looks like this:

  1. Parablocks that have been included by the Inclusion Pipeline are pending approval for a time-window known as the secondary checking window.
  2. During the secondary-checking window, validators randomly self-select to perform secondary checks on the parablock.
  3. These validators, known in this context as secondary checkers, acquire the parablock and its PoV, and re-run the validation function.
  4. The secondary checkers gossip the result of their checks. Contradictory results lead to escalation, where all validators are required to check the block. The validators on the losing side of the dispute are slashed.
  5. At the end of the Approval Process, the parablock is either Approved or it is rejected. More on the rejection process later.

More information on the Approval Process can be found in the dedicated section on Approval. More information on Disputes can be found in the dedicated section on Disputes.

These two pipelines sum up the sequence of events necessary to extend and acquire full security on a Parablock. Note that the Inclusion Pipeline must conclude for a specific parachain before a new block can be accepted on that parachain. After inclusion, the Approval Process kicks off, and can be running for many parachain blocks at once.

Reiterating the lifecycle of a candidate:

  1. Candidate: put forward by a collator to a validator.
  2. Seconded: put forward by a validator to other validators
  3. Backable: validity attested to by a majority of assigned validators
  4. Backed: Backable & noted in a fork of the relay-chain.
  5. Pending availability: Backed but not yet considered available.
  6. Included: Backed and considered available.
  7. Accepted: Backed, available, and undisputed
%3cluster_vgParachain Validators(subset of all)cluster_approvalSecondary Checkers(subset of all)v1Validator 1v2Validator 2v1->v2(2) Secondedv3Validator 3v1->v3(2) Secondedb(3) Backablev1->bv2->bv3->bv4Validator 4 (relay chain)(selected by BABE)b->v4(4) Backedpa(5) Relay Block (Pending Availability)v4->pacolCollatorcol->v1(1) CandidatepbParablockpa->pb(6) a few blocks later...a5Validator 5pb->a5a6Validator 6pb->a6a7Validator 7pb->a7rcRelay Chain Validatorsa5->rc(7) Approveda6->rc(7) Approveda7->rc(7) Approved

The diagram above shows the happy path of a block from (1) Candidate to the (7) Approved state.

It is also important to take note of the fact that the relay-chain is extended by BABE, which is a forkful algorithm. That means that different block authors can be chosen at the same time, and may not be building on the same block parent. Furthermore, the set of validators is not fixed, nor is the set of parachains. And even with the same set of validators and parachains, the validators' assignments to parachains is flexible. This means that the architecture proposed in the next chapters must deal with the variability and multiplicity of the network state.

%3rcaRelay Block ArcbRelay Block Brcb->rcarccRelay Block Crcc->rcbvg1Validator Group 1(Validator 4)(Validator 1) (Validator 2)(Validator 5)vg1->rccBuilding on Cvg2Validator Group 2(Validator 7)(Validator 3) (Validator 6)vg2->rcbBuilding on B

In this example, group 1 has received block C while the others have not due to network asynchrony. Now, a validator from group 2 may be able to build another block on top of B, called C'. Assume that afterwards, some validators become aware of both C and C', while others remain only aware of one.

%3rcaRelay Block ArcbRelay Block Brcb->rcarccRelay Block Crcc->rcbrcc_primeRelay Block C'rcc_prime->rcbvg1Validator Group 1(Validator 4) (Validator 1)vg1->rccvg2Validator Group 2(Validator 7) (Validator 6)vg2->rcc_primevg3Validator Group 3(Validator 2) (Validator 3)(Validator 5)vg3->rccvg3->rcc_prime

Those validators that are aware of many competing heads must be aware of the work happening on each one. They may contribute to some or a full extent on both. It is possible that due to network asynchrony two forks may grow in parallel for some time, although in the absence of an adversarial network this is unlikely in the case where there are validators who are aware of both chain heads.

Approval Process

The Approval Process is the mechanism by which the relay-chain ensures that only valid parablocks are finalized and that backing validators are held accountable for managing to get bad blocks included into the relay chain.

Having a parachain include a bad block into a fork of the relay-chain is not catastrophic as long as the block isn't finalized by the relay-chain's finality gadget, GRANDPA. If the block isn't finalized, that means that the fork of the relay-chain can be reverted in favor of another by means of a dynamic fork-choice rule which leads honest validators to ignore any forks containing that parablock.

Dealing with a bad parablock proceeds in these stages:

  1. Detection
  2. Escalation
  3. Consequences

First, the bad block must be detected by an honest party. Second, the honest party must escalate the bad block to be checked by all validators. And last, the correct consequences of a bad block must occur. The first consequence, as mentioned above, is to revert the chain so what full nodes perceive to be best no longer contains the bad parablock. The second consequence is to slash all malicious validators. Note that, if the chain containing the bad block is reverted, that the result of the dispute needs to be transplanted or at least transplantable to all other forks of the chain so that malicious validators are slashed in all possible histories. Phrased alternatively, there needs to be no possible relay-chain in which malicious validators get away cost-free.

Accepting a parablock is the end result of having passed through the detection stage without dispute, or having passed through the escalation/dispute stage with a positive outcome. For this to work, we need the detection procedure to have the properties that enough honest validators are always selected to check the parablock and that they cannot be interfered with by an adversary. This needs to be balanced with the scaling concern of parachains in general: the easiest way to get the first property is to have everyone check everything, but that is clearly too heavy. So we also have a desired constraint on the other property that we have as few validators as possible check any particular parablock. Our assignment function is the method by which we select validators to do approval checks on parablocks.

It often makes more sense to think of relay-chain blocks as having been approved or not as opposed to thinking about whether parablocks have been approved. A relay-chain block containing a single bad parablock needs to be reverted, and a relay-chain block that contains only approved parablocks can be called approved, as long as its parent relay-chain block is also approved. It is important that the validity of any particular relay-chain block depend on the validity of its ancestry, so we do not finalize a block which has a bad block in its ancestry.

%3IncludedIncludedAssignmentsAssignmentsIncluded->AssignmentsApprovalApprovalAssignments->ApprovalEscalationEscalationAssignments->EscalationFinalityFinalityApproval->FinalityConsequencesConsequencesEscalation->Consequences

Approval has roughly two parts:

  • Assignments determines which validators performs approval checks on which candidates. It ensures that each candidate receives enough random checkers, while reducing adversaries' odds for obtaining enough checkers, and limiting adversaries' foreknowledge. It tracks approval votes to identify when "no show" approval check takes suspiciously long, perhaps indicating the node being under attack, and assigns more checks in this case. It tracks relay chain equivocations to determine when adversaries possibly gained foreknowledge about assignments, and adds additional checks in this case.

  • Approval checks listens to the assignments subsystem for outgoing assignment notices that we shall check specific candidates. It then performs these checks by first invoking the reconstruction subsystem to obtain the candidate, second invoking the candidate validity utility subsystem upon the candidate, and finally sending out an approval vote, or perhaps initiating a dispute.

These both run first as off-chain consensus protocols using messages gossiped among all validators, and second as an on-chain record of this off-chain protocols' progress after the fact. We need the on-chain protocol to provide rewards for the off-chain protocol.

Approval requires two gossiped message types, assignment notices created by its assignments subsystem, and approval votes sent by our approval checks subsystem when authorized by the candidate validity utility subsystem.

Approval keys

We need two separate keys for the approval subsystem:

  • Approval assignment keys are sr25519/schnorrkel keys used only for the assignment criteria VRFs. We implicitly sign assignment notices with approval assignment keys by including their relay chain context and additional data in the VRF's extra message, but exclude these from its VRF input.

  • Approval vote keys would only sign off on candidate parablock validity and has no natural key type restrictions. There's no need for this to actually embody a new session key type. We just want to make a distinction between assignments and approvals, although distant future node configurations might favor separate roles. We re-use the same keys as are used for parachain backing in practice.

Approval vote keys could relatively easily be handled by some hardened signer tooling, perhaps even HSMs assuming we select ed25519 for approval vote keys. Approval assignment keys might or might not support hardened signer tooling, but doing so sounds far more complex. In fact, assignment keys determine only VRF outputs that determine approval checker assignments, for which they can only act or not act, so they cannot equivocate, lie, etc. and represent little if any slashing risk for validator operators.

In future, we shall determine which among the several hardening techniques best benefits the network as a whole. We could provide a multi-process multi-machine architecture for validators, perhaps even reminiscent of GNUNet, or perhaps more resembling smart HSM tooling. We might instead design a system that more resembled full systems, like like Cosmos' sentry nodes. In either case, approval assignments might be handled by a slightly hardened machine, but not necessarily nearly as hardened as approval votes, but approval votes machines must similarly run foreign WASM code, which increases their risk, so assignments being separate sounds helpful.

Assignments

Approval assignment determines on which candidate parachain blocks each validator performs approval checks. An approval session considers only one relay chain block and assigns only those candidates that relay chain block declares available.

Assignment balances several concerns:

  • limits adversaries' foreknowledge about assignments,
  • ensures enough checkers, and
  • distributes assignments relatively equitably.

Assignees determine their own assignments to check specific candidates using two or three assignment criteria. Assignees never reveal their assignments until relevant, and gossip delays assignments sent early, which limits others' foreknowledge. Assignees learn their assignment only with the relay chain block.

All criteria require the validator evaluate a verifiable random function (VRF) using their VRF secret key. All criteria input specific data called "stories" about the session's relay chain block, and output candidates to check and a precedence called a DelayTranche.

We liberate availability cores when their candidate becomes available of course, but one approval assignment criteria continues associating each candidate with the core number it occupied when it became available.

Assignment operates in loosely timed rounds determined by this DelayTranches, which proceed roughly 12 times faster than six second block production assuming half second gossip times. If a candidate C needs more approval checkers by the time we reach round t then any validators with an assignment to C in delay tranche t gossip their send assignment notice for C. We continue until all candidates have enough approval checkers assigned. We take entire tranches together if we do not yet have enough, so we expect strictly more than enough checkers. We also take later tranches if some checkers return their approval votes too slow (see no shows below).

Assignment ensures validators check those relay chain blocks for which they have delay tranche zero aka the highest precedence, so that adversaries always face honest checkers equal to the expected number of assignments with delay tranche zero.

Among these criteria, the BABE VRF output provides the story for two, which reduces how frequently adversaries could position their own checkers. We have one criterion whose story consists of the candidate's block hash plus external knowledge that a relay chain equivocation exists with a conflicting candidate. It provides unforeseeable assignments when adversaries gain foreknowledge about the other two by committing an equivocation in relay chain block production.

Announcements / Notices

We gossip assignment notices among nodes so that all validators know which validators should check each candidate, and if any candidate requires more checkers.

Assignment notices consist of a relay chain context given by a block hash, an assignment criteria, consisting of the criteria identifier and optionally a criteria specific field, an assignee identifier, and a VRF signature by the assignee, which itself consists of a VRF pre-output and a DLEQ proof. Its VRF input consists of the criteria, usually including a criteria specific field, and a "story" about its relay chain context block.

We never include stories inside the gossip messages containing assignment notices, but require each validator reconstruct them. We never care about assignments in the disputes process, so this does not complicate remote disputes.

In a Schnorr VRF, there is an extra signed message distinct from this input, which we set to the relay chain block hash. As a result, assignment notices are self signing and can be "politely" gossiped without additional signatures, meaning between nodes who can compute the story from the relay chain context. In other words, if we cannot compute the story required by an assignment notice's VRF part then our self signing property fails and we cannot verify its origin. We could fix this with either another signature layer (64 bytes) or by including the VRF input point computed from the story (32 bytes), but doing so appears unhelpful.

Any validator could send their assignment notices and/or approval votes too early. We gossip the approval votes early because they represent a major commitment by the validator. We delay gossiping the assignment notices until they agree with our local clock however. We also impose a politeness condition that the recipient knows the relay chain context used by the assignment notice.

Stories

We based assignment criteria upon two possible "stories" about the relay chain block R that included the candidate aka declared the candidate available. All stories have an output that attempts to minimize adversarial influence, which then acts as the VRF input for an assignment criteria.

We first have a RelayVRFStory that outputs the randomness from another VRF output produced by the relay chain block producer when creating R. Among honest nodes, only this one relay chain block producer who creates R knew the story in advance, and even they knew nothing two epochs previously.

In BABE, we create this value calling schnorrkel::vrf::VRFInOut::make_bytes with a context "A&V RC-VRF", with the VRFInOut coming from either the VRF that authorized block production for primary blocks, or else from the secondary block VRF for the secondary block type.

In Sassafras, we shall always use the non-anonymized recycling VRF output, never the anonymized ring VRF that authorizes block production. We do not currently know if Sassafras shall have a separate schnorrkel key, but if it reuses its ring VRF key there is an equivalent ring_vrf::VRFInOut::make_bytes.

We like that RelayVRFStory admits relatively few choices, but an adversary who equivocates in relay chain block production could learn assignments that depend upon the RelayVRFStory too early because the same relay chain VRF appears in multiple blocks.

We therefore provide a secondary RelayEquivocationStory that outputs the candidate's block hash, but only for candidate equivocations. We say a candidate C in R is an equivocation when there exists another relay chain block R1 that equivocates for R in the sense that R and R1 have the same RelayVRFStory, but R contains C and R1 does not contain C.

We want checkers for candidate equivocations that lie outside our preferred relay chain as well, which represents a slightly different usage for the assignments module, and might require more information in the gossip messages.

Assignment criteria

Assignment criteria compute actual assignments using stories and the validators' secret approval assignment key. Assignment criteria output a Position consisting of both a ParaId to be checked, as well as a precedence DelayTranche for when the assignment becomes valid.

Assignment criteria come in three flavors, RelayVRFModulo, RelayVRFDelay and RelayEquivocation. Among these, both RelayVRFModulo and RelayVRFDelay run a VRF whose input is the output of a RelayVRFStory, while RelayEquivocation runs a VRF whose input is the output of a RelayEquivocationStory.

Among these, we have two distinct VRF output computations:

RelayVRFModulo runs several distinct samples whose VRF input is the RelayVRFStory and the sample number. It computes the VRF output with schnorrkel::vrf::VRFInOut::make_bytes using the context "A&V Core", reduces this number modulo the number of availability cores, and outputs the candidate just declared available by, and included by aka leaving, that availability core. We drop any samples that return no candidate because no candidate was leaving the sampled availability core in this relay chain block. We choose three samples initially, but we could make polkadot more secure and efficient by increasing this to four or five, and reducing the backing checks accordingly. All successful RelayVRFModulo samples are assigned delay tranche zero.

There is no sampling process for RelayVRFDelay and RelayEquivocation. We instead run them on specific candidates and they compute a delay from their VRF output. RelayVRFDelay runs for all candidates included under, aka declared available by, a relay chain block, and inputs the associated VRF output via RelayVRFStory. RelayEquivocation runs only on candidate block equivocations, and inputs their block hashes via the RelayEquivocation story.

RelayVRFDelay and RelayEquivocation both compute their output with schnorrkel::vrf::VRFInOut::make_bytes using the context "A&V Tranche" and reduce the result modulo num_delay_tranches + zeroth_delay_tranche_width, and consolidate results 0 through zeroth_delay_tranche_width to be 0. In this way, they ensure the zeroth delay tranche has zeroth_delay_tranche_width+1 times as many assignments as any other tranche.

As future work (or TODO?), we should merge assignment notices with the same delay and story using vrf_merge. We cannot merge those with the same delay and different stories because RelayEquivocationStorys could change but RelayVRFStory never changes.

Announcer and Watcher/Tracker

We track all validators' announced approval assignments for each candidate associated to each relay chain block, which tells us which validators were assigned to which candidates.

We permit at most one assignment per candidate per story per validator, so one validator could be assigned under both the RelayVRFDelay and RelayEquivocation criteria, but not under both RelayVRFModulo and RelayVRFDelay criteria, since those both use the same story. We permit only one approval vote per candidate per validator, which counts for any applicable criteria.

We announce, and start checking for, our own assignments when the delay of their tranche is reached, but only if the tracker says the assignee candidate requires more approval checkers. We never announce an assignment we believe unnecessary because early announcements gives an adversary information. All delay tranche zero assignments always get announced, which includes all RelayVRFModulo assignments.

In other words, if some candidate C needs more approval checkers by the time we reach round t then any validators with an assignment to C in delay tranche t gossip their send assignment notice for C, and begin reconstruction and validation for 'C. If however C reached enough assignments, then validators with later assignments skip announcing their assignments.

We continue until all candidates have enough approval checkers assigned. We never prioritize assignments within tranches and count all or no assignments for a given tranche together, so we often overshoot the target number of assigned approval checkers.

No shows

We have a "no show" timeout longer than one relay chain slot, so at least 6 seconds, during which we expect approval checks should succeed in reconstructing the candidate block, in redoing its erasure coding to check the candidate receipt, and finally in rechecking the candidate block itself.

We consider a validator a "no show" if they do not approve or dispute within this "no show" timeout from our receiving their assignment notice. We time this from our receipt of their assignment notice instead of our imagined real time for their tranche because otherwise receiving late assignment notices creates immediate "no shows" and unnecessary work.

We worry "no shows" represent a validator under denial of service attack, presumably to prevent it from reconstructing the candidate, but perhaps delaying it form gossiping a dispute too. We therefore always replace "no shows" by adding one entire extra delay tranche worth of validators, so such attacks always result in additional checkers.

As an example, imagine we need 20 checkers, but tranche zero produces only 14, and tranche one only 4, then we take all 5 from tranche two, and thus require 23 checkers for that candidate. If one checker Charlie from tranche one or two does not respond within say 8 seconds, then we add all 7 checkers from tranche three. If again one checker Cindy from tranche three does not respond within 8 seconds then we take all 3 checkers from tranche four. We now have 33 checkers working on the candidate, so this escalated quickly.

We escalated so quickly because we worried that Charlie and Cindy might be the only honest checkers assigned to that candidate. If therefore either Charlie or Cindy finally return an approval, then we can conclude approval, and abandon the checkers from tranche four.

We therefore require the "no show" timeout to be longer than a relay chain slot so that we can witness "no shows" on-chain. We discuss below how this helps reward validators who replace "no shows".

We avoid slashing for "no shows" by itself, although being "no show" could enter into some computation that punishes repeated poor performance, presumably replaces ImOnline, and we could reduce their rewards and further rewards those who filled in.

As future work, we foresee expanding the "no show" scheme to anonymize the additional checkers, like by using assignment noticed with a new criteria that employs a ring VRF and then all validators providing cover by requesting a couple erasure coded pieces, but such anonymity scheme sound extremely complex and lie far beyond our initial functionality.

Assignment postponement

We expect validators could occasionally overloaded when they randomly acquire too many assignments. All these fluctuations amortize over multiple blocks fairly well, but this slows down finality.

We therefore permit validators to delay sending their assignment noticed intentionally. If nobody knows about their assignment then they avoid creating "no shows" and the workload progresses normally.

We strongly prefer if postponements come from tranches higher aka less important than zero because tranche zero checks provide somewhat more security.

TODO: When? Is this optimal for the network? etc.

On-chain verification

We should verify approval on-chain to reward approval checkers. We therefore require the "no show" timeout to be longer than a relay chain slot so that we can witness "no shows" on-chain, which helps with this goal. The major challenge with an on-chain record of the off-chain process is adversarial block producers who may either censor votes or publish votes to the chain which cause other votes to be ignored and unrewarded (reward stealing).

In principle, all validators have some "tranche" at which they're assigned to the parachain candidate, which ensures we reach enough validators eventually. As noted above, we often retract "no shows" when the slow validator eventually shows up, so witnessing their initially being a "no show" helps manage rewards.

We expect on-chain verification should work in two phases: We first record assignments notices and approval votes on-chain in relay chain block, doing the VRF or regular signature verification again in block verification, and inserting chain authenticated unsigned notes into the relay chain state that contain the checker, tranche, paraid, and relay block height for each assignment notice. We then later have another relay chain block that runs some "approved" intrinsic, which extract all these notes from the state and feeds them into our approval code.

We now encounter one niche concern in the interaction between postponement and on-chain verification: Any validator with a tranche zero (or other low) assignment could delay sending an assignment notice, like because they postponed their assigned tranche (which is allowed). If they later send this assignment notices right around finality time, then they race with this approved. intrinsic: If their announcement gets on-chain (also allowed), then yes it delays finality. If it does not get on-chain, then yes we've one announcement that the off-chain consensus system says is valid, but the chain ignores for being too slow.

We need the chain to win in this case, but doing this requires imposing an annoyingly long overarching delay upon finality. We might explore limits on postponement too, but this sounds much harder.

Parameters

We prefer doing approval checkers assignments under RelayVRFModulo as opposed to RelayVRFDelay because RelayVRFModulo avoids giving individual checkers too many assignments and tranche zero assignments benefit security the most. We suggest assigning at least 16 checkers under RelayVRFModulo although assignment levels have never been properly analyzed.

Our delay criteria RelayVRFDelay and RelayEquivocation both have two primary paramaters, expected checkers per tranche and the zeroth delay tranche width.

We require expected checkers per tranche to be less than three because otherwise an adversary with 1/3 stake could force all nodes into checking all blocks. We strongly recommend expected checkers per tranche to be less than two, which helps avoid both accidental and intentional explosions. We also suggest expected checkers per tranche be larger than one, which helps prevent adversaries from predicting than advancing one tranche adds only their own validators.

We improve security more with tranche zero assignments, so RelayEquivocation should consolidates its first several tranches into tranche zero. We describe this as the zeroth delay tranche width, which initially we set to 12 for RelayEquivocation and 1 for RelayVRFDelay.

Why VRFs?

We do assignments with VRFs to give "enough" checkers some meaning beyond merely "expected" checkers:

We could specify a protocol that used only system randomness, which works because our strongest defense is the expected number of honest checkers who assign themselves. In this, adversaries could trivially flood their own blocks with their own checkers, so this strong defense becomes our only defense, and delay tranches become useless, so some blocks actually have zero approval checkers and possibly only one checker overall.

VRFs though require adversaries wait far longer between such attacks, which also helps against adversaries with little at stake because they compromised validators. VRFs raise user confidence that no such "drive by" attacks occurred because the delay tranche system ensure at least some minimum number of approval checkers. In this vein, VRFs permit reducing backing checks and increasing approval checks, which makes polkadot more efficient.

Gossip

Any validator could send their assignment notices and/or approval votes too early. We gossip the approval votes because they represent a major commitment by the validator. We retain but delay gossiping the assignment notices until they agree with our local clock.

Assignment notices being gossiped too early might create a denial of service vector. If so, we might exploit the relative time scheme that synchronizes our clocks, which conceivably permits just dropping excessively early assignments.

Finality GRANDPA Voting Rule

The relay-chain requires validators to participate in GRANDPA. In GRANDPA, validators submit off-chain votes on what they believe to be the best block of the chain, and GRANDPA determines the common block contained by a supermajority of sub-chains. There are also additional constraints on what can be submitted based on results of previous rounds of voting.

In order to avoid finalizing anything which has not received enough approval votes or is disputed, we will pair the approval protocol with an alteration to the GRANDPA voting strategy for honest nodes which causes them to vote only on chains where every parachain candidate within has been approved. Furthermore, the voting rule prevents voting for chains where there is any live dispute or any dispute has resolved to a candidate being invalid.

Thus, the finalized relay-chain should contain only relay-chain blocks where a majority believe that every block within has been sufficiently approved.

Future work

We could consider additional gossip messages with which nodes claims "slow availability" and/or "slow candidate" to fine tune the assignments "no show" system, but long enough "no show" delays suffice probably.

We shall develop more practical experience with UDP once the availability system works using direct UDP connections. In this, we should discover if reconstruction performs adequately with a complete graphs or benefits from topology restrictions. At this point, an assignment notices could implicitly request pieces from a random 1/3rd, perhaps topology restricted, which saves one gossip round. If this preliminary fast reconstruction fails, then nodes' request alternative pieces directly. There is an interesting design space in how this overlaps with "slow availability" claims.

Disputes

Fast forward to more detailed disputes requirements.

Motivation and Background

All parachain blocks that end up in the finalized relay chain should be valid. This does not apply to blocks that are only backed, but not included.

We have two primary components for ensuring that nothing invalid ends up in the finalized relay chain:

  • Approval Checking, as described here and implemented according to the Approval Voting subsystem. This protocol can be shown to prevent invalid parachain blocks from making their way into the finalized relay chain as long as the amount of attempts are limited.
  • Disputes, this protocol, which ensures that each attempt to include something bad is caught, and the offending validators are punished. Disputes differ from backing and approval process (and can not be part of those) in that a dispute is independent of a particular fork, while both backing and approval operate on particular forks. This distinction is important! Approval voting stops, if an alternative fork which might not contain the currently approved candidate gets finalized. This is totally fine from the perspective of approval voting as its sole purpose is to make sure invalid blocks won't get finalized. For disputes on the other hand we have different requirements: Even though the "danger" is past and the adversaries were not able to get their invalid block approved, we still want them to get slashed for the attempt. Otherwise they just have been able to get a free try, but this is something we need to avoid in our security model, as it is based on the assumption that the probability of getting an invalid block finalized is very low and an attacker would get bankrupt before it could have tried often enough.

Every dispute stems from a disagreement among two or more validators. If a bad actor creates a bad block, but the bad actor never distributes it to honest validators, then nobody will dispute it. Of course, such a situation is not even an attack on the network, so we don't need to worry about defending against it.

We are interested in identifying and deterring the following attack scenario:

  • A parablock included on a branch of the relay chain is bad

We are also interested in identifying these additional scenarios:

  • A parablock backed on a branch of the relay chain is bad
  • A parablock seconded, but not backed on any branch of the relay chain, is bad.

Punishing misbehavior in the latter two scenarios doesn't effect our security guarantees and introduces substantial technical challenges as described in the No Disputes for Non Included Candidates section of Dispute Coordinator. We therefore choose to punt on disputes in these cases, instead favoring the protocol simplicity resulting from only punishing in the first scenario.

As covered in the protocol overview, checking a parachain block requires 3 pieces of data: the parachain validation code, the AvailableData, and the CandidateReceipt. The validation code is available on-chain, and published ahead of time, so that no two branches of the relay chain have diverging views of the validation code for a given parachain. Note that only for the first scenario, where the parablock has been included on a branch of the relay chain, is the data necessarily available. Thus, dispute processes should begin with an availability process to ensure availability of the AvailableData. This availability process will conclude quickly if the data is already available. If the data is not already available, then the initiator of the dispute must make it available.

Disputes have both an on-chain and an off-chain component. Slashing and punishment is handled on-chain, so votes by validators on either side of the dispute must be placed on-chain. Furthermore, a dispute on one branch of the relay chain should be transposed to all other active branches of the relay chain. The fact that slashing occurs in all histories is crucial for deterring attempts to attack the network. The attacker should not be able to escape with their funds because the network has moved on to another branch of the relay chain where no attack was attempted.

In fact, this is why we introduce a distinction between local and remote disputes. We categorize disputes as either local or remote relative to any particular branch of the relay chain. Local disputes are about dealing with our first scenario, where a parablock has been included on the specific branch we are looking at. In these cases, the chain is corrupted all the way back to the point where the parablock was backed and must be discarded. However, as mentioned before, the dispute must propagate to all other branches of the relay chain. All other disputes are considered remote. For the on-chain component, when handling a dispute for a block which was not included in the current fork of the relay chain, it is impossible to discern between our attack scenarios. It is possible that the parablock was included somewhere, or backed somewhere, or wasn't backed anywhere. The on-chain component for handling these cases will be the same.

Initiation

Disputes are initiated by any validator who finds their opinion on the validity of a parablock in opposition to another issued statement. As all statements currently gathered by the relay chain imply validity, disputes will be initiated only by nodes which perceive that the parablock is bad.

The initiation of a dispute begins off-chain. A validator signs a message indicating that it disputes the validity of the parablock and notifies all other validators, off-chain, of all of the statements it is aware of for the disputed parablock. These may be backing statements or approval-checking statements. It is worth noting that there is no special message type for initiating a dispute. It is the same message as is used to participate in a dispute and vote negatively. As such, there is no consensus required on who initiated a dispute, only on the fact that there is a dispute in-progress.

In practice, the initiator of a dispute will be either one of the backers or one of the approval checkers for the parablock. If the result of execution is found to be invalid, the validator will initiate the dispute as described above. Furthermore, if the dispute occurs during the backing phase, the initiator must make the data available to other validators. If the dispute occurs during approval checking, the data is already available.

Lastly, it is possible that for backing disputes, i.e. where the data is not already available among all validators, that an adversary may DoS the few parties who are checking the block to prevent them from distributing the data to other validators participating in the dispute process. Note that this can only occur pre-inclusion for any given parablock, so the downside of this attack is small and it is not security-critical to address these cases. However, we assume that the adversary can only prevent the validator from issuing messages for a limited amount of time. We also assume that there is a side-channel where the relay chain's governance mechanisms can trigger disputes by providing the full PoV and candidate receipt on-chain manually.

Dispute Participation

Once becoming aware of a dispute, it is the responsibility of all validators to participate in the dispute. Concretely, this means:

  • Circulate all statements about the candidate that we are aware of - backing statements, approval checking statements, and dispute statements.
  • If we have already issued any type of statement about the candidate, go no further.
  • Download the AvailableData. If possible, this should first be attempted from other dispute participants or backing validators, and then (via erasure-coding) from all validators.
  • Extract the Validation Code from any recent relay chain block. Code is guaranteed to be kept available on-chain, so we don't need to download any particular fork of the chain.
  • Execute the block under the validation code, using the AvailableData, and check that all outputs are correct, including the erasure-root of the CandidateReceipt.
  • Issue a dispute participation statement to the effect of the validity of the candidate block.

Disputes conclude after ⅔ supermajority is reached in either direction.

The on-chain component of disputes can be initiated by providing any two conflicting votes and it also waits for a ⅔ supermajority on either side. The on-chain component also tracks which parablocks have already been disputed so the same parablock may only be disputed once on any particular branch of the relay chain. Lastly, it also tracks which blocks have been included on the current branch of the relay chain. When a dispute is initiated for a para, inclusion is halted for the para until the dispute concludes.

The author of a relay chain block should initiate the on-chain component of disputes for all disputes which the chain is not aware of, and provide all statements to the on-chain component as well. This should all be done via inherents.

Validators can learn about dispute statements in two ways:

  • Receiving them from other validators over gossip
  • Scraping them from imported blocks of the relay chain. This is also used for validators to track other types of statements, such as backing statements.

Validators are rewarded for providing statements to the chain as well as for participating in the dispute, on either side. However, the losing side of the dispute is slashed.

Dispute Conclusion

Disputes, roughly, are over when one side reaches a ⅔ supermajority. They may also never conclude without either side witnessing supermajority, which will only happen if the majority of validators are unable to vote for some reason. Furthermore, disputes on-chain will stay open for some fixed amount of time even after concluding, to accept new votes.

Late votes, after the dispute already reached a ⅔ supermajority, must be rewarded (albeit a smaller amount) as well.

Disputes Flows

A component-free description in what-if form with addition state graphs of the dispute.

stateDiagram-v2
    [*] --> WaitForBackingVote: negative Vote received
    [*] --> WaitForDisputeVote: backing Vote received
    WaitForBackingVote --> Open: negative Vote received
    WaitForDisputeVote --> Open: backing Vote received
    Open --> Concluded: Incoming Vote via Gossip
    Open --> Open: No ⅔ supermajority
    Open --> [*]
    Concluded --> [*]

stateDiagram-v2
    [*] --> Open: First Vote(s) received
    Open --> HasPoV : Fetch Availability Store for PoV

    HasPoV --> HasCode : Fetch historical Code
    HasCode --> VerifyWithRuntime: All Data locally avail

    Open --> DisputeAvailabilityDataReceived
    DisputeAvailabilityDataReceived --> VerifyWithRuntime: Received Gossip

    HasPoV --> RequestDisputeAvailabilityData: nope
    HasCode --> RequestDisputeAvailabilityData: nope
    RequestDisputeAvailabilityData --> VerifyWithRuntime: Received
    RequestDisputeAvailabilityData --> RequestDisputeAvailabilityData: Timed out - pick another peer

    VerifyWithRuntime --> CastVoteValid: Block Valid
    VerifyWithRuntime --> CastVoteInvalid: Block Invalid
    CastVoteInvalid --> GossipVote
    CastVoteValid --> GossipVote
    GossipVote --> [*]


Dispute Availability Data

stateDiagram-v2
    [*] --> Open: First Vote(s) received
    Open --> DisputeDataAvail: somehow the data became available
    Open --> RespondUnavailable: Data not available
    IncomingRequestDisputeAvailabilityData --> RespondUnavailable
    IncomingRequestDisputeAvailabilityData --> DisputeDataAvail
    DisputeDataAvail --> RespondWithDisputeAvailabilityData: Send
    VoteGossipReceived --> Track: implies source peer has<br />dispute availablity data

Peer handling

stateDiagram-v2
    [*] --> Open: First Vote(s) received
    Open --> GossipVotes: for all current peers
    Open --> PeerConnected: another
    PeerConnected --> GossipVotes: Peer connects
    GossipVotes --> [*]

Conditional formulation

The set of validators eligible to vote consists of the validators that had duty at the time of backing, plus backing votes by the backing validators.

If a validator receives an initial dispute message (a set of votes where there are at least two opposing votes contained), and the PoV or Code are hence not reconstructable from local storage, that validator must request the required data from its peers.

The dispute availability message must contain code, persisted validation data, and the proof of validity.

Only peers that already voted shall be queried for the dispute availability data.

The peer to be queried for disputes data, must be picked at random.

A validator must retain code, persisted validation data and PoV until a block, that contains the dispute resolution, is finalized - plus an additional 24 hours.

Dispute availability gossip must continue beyond the dispute resolution, until the post resolution timeout expired (equiv to the timeout until which additional late votes are accepted).

Remote disputes are disputes that are in relation to a chain that is not part of the local validators active heads.

All incoming votes must be persisted.

Persisted votes stay persisted for N sessions, and are cleaned up on a per session basis.

Votes must be queryable by a particular validator, identified by its signing key.

Votes must be queryable by a particular validator, identified by a session index and the validator index valid in that session.

If there exists a negative and a positive vote for a particular block, a dispute is detected.

If a dispute is detected, all currently available votes for that block must be gossiped.

If an incoming dispute vote is detected, a validator must cast their own vote. The vote is determined by validating the PoV with the Code at the time of backing the block in question.

If the validator was also a backer of the block, validation and casting an additional vote should be skipped.

If the count of votes pro or cons regarding the disputed block, reaches the required ⅔ supermajority (including the backing votes), the conclusion must be recorded on chain and the voters on the loosing and no-shows being slashed appropriately.

If a block is found invalid by a dispute resolution, it must be blacklisted to avoid resync or further build on that chain if other chains are available (to be detailed in the grandpa fork choice rule).

A dispute accepts Votes after the dispute is resolved, for 1 day.

If a vote is received, after the dispute is resolved, the vote shall still be recorded in the state root, albeit yielding less reward.

Recording in the state root might happen batched, at timeout expiry.

If a new active head/chain appears, and the dispute resolution was not recorded on that chain yet, the dispute resolution or open dispute must be recorded / transplanted to that chain as well, since the disputes must be present on all chains to make sure the offender is punished.

If a validator votes in two opposing ways, this composes of a double vote like in other cases (backing, approval voting).

If a dispute is not resolved within due time, all validators are to be slashed for a small amount.

If a dispute is not resolved within due time, governance mode shall be entered for manual resolution.

If a validator unexpectedly restarts, the dispute shall be continued with the state based on votes being cast and being present in persistent storage.

Chain Selection

Chain selection processes in blockchains are used for the purpose of selecting blocks to build on and finalize. It is important for these processes to be consistent among nodes and resilient to a maximum proportion of malicious nodes which do not obey the chain selection process.

The parachain host uses both a block authoring system and a finality gadget. The chain selection strategy of the parachain host involves two key components: a leaf-selection rule and a set of finality constraints. When it's a validator's turn to author on a block, they are expected to select the best block via the leaf-selection rule to build on top of. When a validator is participating in finality, there is a minimum block which can be voted on, which is usually the finalized block. The validator should select the best chain according to the leaf-selection rule and subsequently apply the finality constraints to arrive at the actual vote cast by that validator.

Before diving into the particularities of the leaf-selection rule and the finality constraints, it's important to discuss the goals that these components are meant to achieve. For this it is useful to create the definitions of viable and finalizable blocks.

Property Definitions

A block is considered viable when all of the following hold:

  1. It is or descends from the finalized block
  2. It is not stagnant
  3. It is not reverted.

A block is considered a viable leaf when all of the following hold:

  1. It is viable
  2. It has no viable descendant.

A block is considered stagnant when either:

  1. It is unfinalized, is not approved, and has not been approved within 2 minutes
  2. Its parent is stagnant.

A block is considered reverted when either:

  1. It is unfinalized and includes a candidate which has lost a dispute
  2. Its parent is reverted

A block is considered finalizable when all of the following hold:

  1. It is viable
  2. Its parent, if unfinalized, is finalizable.
  3. It is either finalized or approved.
  4. It is either finalized or includes no candidates which have unresolved disputes or have lost a dispute.

The leaf-selection rule

We assume that every block has an implicit weight or score which can be used to compare blocks. In BABE, this is determined by the number of primary slots included in the chain. In PoW, this is the chain with either the most work or GHOST weight.

The leaf-selection rule based on our definitions above is simple: we take the maximum-scoring viable leaf we are aware of. In the case of a tie we select the one with a lower lexicographical block hash.

The best-chain-containing rule

Finality gadgets, as mentioned above, will often impose an additional requirement to vote on a chain containing a specific block, known as the required block. Although this is typically the most recently finalized block, it is possible that it may be a block that is unfinalized. When receiving such a request:

  1. If the required block is the best finalized block, then select the best viable leaf.
  2. If the required block is unfinalized and non-viable, then select the required block and go no further. This is likely an indication that something bad will be finalized in the network, which will never happen when approvals & disputes are functioning correctly. Nevertheless we account for the case here.
  3. If the required block is unfinalized and viable, then iterate over the viable leaves in descending order by score and select the first one which contains the required block in its chain. Backwards iteration is a simple way to check this, but if unfinalized chains grow long then Merkle Mountain-Ranges will most likely be more efficient.

Once selecting a leaf, the chain should be constrained to the maximum of the required block or the highest finalizable ancestor.

Architecture Overview

This section aims to describe, at a high level, the code architecture and subsystems involved in the implementation of an individual Parachain Host. It also illuminates certain subtleties and challenges faced in the design and implementation of those subsystems.

To recap, Polkadot includes a blockchain known as the relay-chain. A blockchain is a Directed Acyclic Graph (DAG) of state transitions, where every block can be considered to be the head of a linked-list (known as a "chain" or "fork") with a cumulative state which is determined by applying the state transition of each block in turn. All paths through the DAG terminate at the Genesis Block. In fact, the blockchain is a tree, since each block can have only one parent.

%3genesisGenesisb1Block 1b1->genesisb2Block 2b2->genesisb3Block 3b3->b1b4Block 4b4->b3b5Block 5b5->b3

A blockchain network is comprised of nodes. These nodes each have a view of many different forks of a blockchain and must decide which forks to follow and what actions to take based on the forks of the chain that they are aware of.

So in specifying an architecture to carry out the functionality of a Parachain Host, we have to answer two categories of questions:

  1. What is the state-transition function of the blockchain? What is necessary for a transition to be considered valid, and what information is carried within the implicit state of a block?
  2. Being aware of various forks of the blockchain as well as global private state such as a view of the current time, what behaviors should a node undertake? What information should a node extract from the state of which forks, and how should that information be used?

The first category of questions will be addressed by the Runtime, which defines the state-transition logic of the chain. Runtime logic only has to focus on the perspective of one chain, as each state has only a single parent state.

The second category of questions addressed by Node-side behavior. Node-side behavior defines all activities that a node undertakes, given its view of the blockchain/block-DAG. Node-side behavior can take into account all or many of the forks of the blockchain, and only conditionally undertake certain activities based on which forks it is aware of, as well as the state of the head of those forks.

GRuntimeRuntimeNodeNodeRuntime->NodeRuntime APITransportTransportNode->Transport

It is also helpful to divide Node-side behavior into two further categories: Networking and Core. Networking behaviors relate to how information is distributed between nodes. Core behaviors relate to internal work that a specific node does. These two categories of behavior often interact, but can be heavily abstracted from each other. Core behaviors care that information is distributed and received, but not the internal details of how distribution and receipt function. Networking behaviors act on requests for distribution or fetching of information, but are not concerned with how the information is used afterwards. This allows us to create clean boundaries between Core and Networking activities, improving the modularity of the code.

          ___________________                    ____________________
         /       Core        \                  /     Networking     \
         |                   |  Send "Hello"    |                    |
         |                   |-  to "foo"   --->|                    |
         |                   |                  |                    |
         |                   |                  |                    |
         |                   |                  |                    |
         |                   |    Got "World"   |                    |
         |                   |<--  from "bar" --|                    |
         |                   |                  |                    |
         \___________________/                  \____________________/
                                                   ______| |______
                                                   ___Transport___

Node-side behavior is split up into various subsystems. Subsystems are long-lived workers that perform a particular category of work. Subsystems can communicate with each other, and do so via an Overseer that prevents race conditions.

Runtime logic is divided up into Modules and APIs. Modules encapsulate particular behavior of the system. Modules consist of storage, routines, and entry-points. Routines are invoked by entry points, by other modules, upon block initialization or closing. Routines can read and alter the storage of the module. Entry-points are the means by which new information is introduced to a module and can limit the origins (user, root, parachain) that they accept being called by. Each block in the blockchain contains a set of Extrinsics. Each extrinsic targets a a specific entry point to trigger and which data should be passed to it. Runtime APIs provide a means for Node-side behavior to extract meaningful information from the state of a single fork.

These two aspects of the implementation are heavily dependent on each other. The Runtime depends on Node-side behavior to author blocks, and to include Extrinsics which trigger the correct entry points. The Node-side behavior relies on Runtime APIs to extract information necessary to determine which actions to take.

Messaging Overview

The Polkadot Host has a few mechanisms that are responsible for message passing. They can be generally divided on two categories: Horizontal and Vertical. Horizontal Message Passing (HMP) refers to mechanisms that are responsible for exchanging messages between parachains. Vertical Message Passing (VMP) is used for communication between the relay chain and parachains.

Vertical Message Passing

%3rcRelay Chainp1Parachainrc->p1DMPp1->rcUMP

Downward Message Passing (DMP) is a mechanism for delivering messages to parachains from the relay chain.

Each parachain has its own queue that stores all pending inbound downward messages. A parachain doesn't have to process all messages at once, however, there are rules as to how the downward message queue should be processed. Currently, at least one message must be consumed per candidate if the queue is not empty. The downward message queue doesn't have a cap on its size and it is up to the relay-chain to put mechanisms that prevent spamming in place.

Upward Message Passing (UMP) is a mechanism responsible for delivering messages in the opposite direction: from a parachain up to the relay chain. Upward messages are essentially byte blobs. However, they are interpreted by the relay-chain according to the XCM standard.

The XCM standard is a common vocabulary of messages. The XCM standard doesn't require a particular interpretation of a message. However, the parachains host (e.g. Polkadot) guarantees certain semantics for those.

Moreover, while most XCM messages are handled by the on-chain XCM interpreter, some of the messages are special cased. Specifically, those messages can be checked during the acceptance criteria and thus invalid messages would lead to rejecting the candidate itself.

One kind of such a message is Xcm::Transact. This upward message can be seen as a way for a parachain to execute arbitrary entrypoints on the relay-chain. Xcm::Transact messages resemble regular extrinsics with the exception that they originate from a parachain.

The payload of Xcm::Transact messages is referred as to Dispatchable. When a candidate with such a message is enacted the dispatchables are put into a queue corresponding to the parachain. There can be only so many dispatchables in that queue at once. The weight that processing of the dispatchables can consume is limited by a preconfigured value. Therefore, it is possible that some dispatchables will be left for later blocks. To make the dispatching more fair, the queues are processed turn-by-turn in a round robin fashion.

The second category of special cased XCM messages are for horizontal messaging channel management, namely messages meant to request opening and closing HRMP channels (HRMP will be described below).

Horizontal Message Passing

%3rcRelay Chainp1Parachain 1rc->p1DMPp2Parachain 2rc->p2DMPp1->rcUMPp2->rcUMPp2->p1XCMP

Cross-Chain Message Passing

The most important member of this family is XCMP.

ℹ️ XCMP is currently under construction and details are subject for change.

XCMP is a message passing mechanism between parachains that require minimal involvement of the relay chain. The relay chain provides means for sending parachains to authenticate messages sent to recipient parachains.

Semantically communication occurs through so called channels. A channel is unidirectional and it has two endpoints, for sender and for recipient. A channel can be opened only if the both parties agree and closed unilaterally.

Only the channel metadata is stored on the relay-chain in a very compact form: all messages and their contents sent by the sender parachain are encoded using only one root hash. This root is referred as MQC head.

The authenticity of the messages must be proven using that root hash to the receiving party at the candidate authoring time. The proof stems from the relay parent storage that contains the root hash of the channel. Since not all messages are required to be processed by the receiver's candidate, only the processed messages are supplied (i.e. preimages), rest are provided as hashes.

Further details can be found at the official repository for the Cross-Consensus Message Format (XCM), as well as at the W3F research website and this blogpost.

HRMP (Horizontally Relay-routed Message Passing) is a stop gap that predates XCMP. Semantically, it mimics XCMP's interface. The crucial difference from XCMP though is that all the messages are stored in the relay-chain storage. That makes things simple but at the same time that makes HRMP more demanding in terms of resources thus making it more expensive.

Once XCMP is available we expect to retire HRMP.

PVF Pre-checking Overview

Motivation

Parachains' validation function is described by a wasm module that we refer to as a PVF. Since a PVF is a wasm module the typical way of executing it is to compile it to machine code.

Typically an optimizing compiler consists of algorithms that are able to optimize the resulting machine code heavily. However, while those algorithms perform quite well for a typical wasm code produced by standard toolchains (e.g. rustc/LLVM), those algorithms can be abused to consume a lot of resources. Moreover, since those algorithms are rather complex there is a lot of room for a bug that can crash the compiler.

If compilation of a Parachain Validation Function (PVF) takes too long or uses too much memory, this can leave a node in limbo as to whether a candidate of that parachain is valid or not.

The amount of time that a PVF takes to compile is a subjective resource limit and as such PVFs may be maliciously crafted so that there is e.g. a 50/50 split of validators which can and cannot compile and execute the PVF.

This has the following implications:

  • In backing, inclusion may be slow due to backing groups being unable to execute the block
  • In approval checking, there may be many no-shows, leading to slow finality
  • In disputes, neither side may reach supermajority. Nobody will get slashed and the chain will not be reverted or finalized.

As a result of this issue we need a fairly hard guarantee that the PVFs of registered parachains/threads can be compiled within a reasonable amount of time.

Solution

The problem is solved by having a pre-checking process.

Pre-checking

Pre-checking mostly consists of attempting to prepare (compile) the PVF WASM blob. We use more strict limits (e.g. timeouts) here compared to regular preparation for execution. This way errors during preparation later are likely unrelated to the PVF itself, as it already passed pre-checking. We can treat such errors as local node issues.

We also have an additional step where we attempt to instantiate the WASM runtime without running it. This is unrelated to preparation so we don't time it, but it does help us catch more issues.

Protocol

Pre-checking is run when a new validation code is included in the chain. A new PVF can be added in two cases:

  • A new parachain is registered.
  • An existing parachain signalled an upgrade of its validation code.

Before any of those operations finish, the PVF pre-checking vote is initiated. The PVF pre-checking vote is identified by the PVF code hash that is being voted on. If there is already PVF pre-checking process running, then no new PVF pre-checking vote will be started. Instead, the operation just subscribes to the existing vote.

The pre-checking vote can be concluded either by obtaining a threshold of votes for a decision, or if it expires. The threshold to accept is a supermajority of 2/3 of validators. We reject once a supermajority is no longer possible.

Each validator checks the list of PVFs available for voting. The vote is binary, i.e. accept or reject a given PVF. As soon as the threshold of votes are collected for one of the sides of the vote, the voting is concluded in that direction and the effects of the voting are enacted.

Only validators from the active set can participate in the vote. The set of active validators can change each session. That's why we reset the votes each session. A voting that observed a certain number of sessions will be rejected.

The effects of the PVF accepting depend on the operations requested it:

  1. All onboardings subscribed to the approved PVF pre-checking process will get scheduled and after passing 2 session boundaries they will be onboarded.
  2. All upgrades subscribed to the approved PVF pre-checking process will get scheduled very similarly to the existing process. Upgrades with pre-checking are really the same process that is just delayed by the time required for pre-checking voting. In case of instant approval the mechanism is exactly the same.

In case PVF pre-checking process was concluded with rejection, then all the operations that are subscribed to the rejected PVF pre-checking process will be processed as follows. That is, onboarding or upgrading will be cancelled.

The logic described above is implemented by the paras module.

Subsystem

On the node-side, there is a PVF pre-checking subsystem that scans the chain for new PVFs via using runtime APIs. Upon finding a new PVF, the subsystem will initiate a PVF pre-checking request and wait for the result. Whenever the result is obtained, the subsystem will use the runtime API to submit a vote for the PVF. The vote is an unsigned transaction. The vote will be distributed via the gossip similarly to a normal transaction. Eventually a block producer will include the vote into the block where it will be handled by the runtime.

Summary

Parachains' validation function is described by a wasm module that we refer to as a PVF.

In order to make the PVF usable for candidate validation it has to be registered on-chain.

As part of the registration process, it has to go through pre-checking. Pre-checking is a game of attempting preparation and additional checks, and reporting the results back on-chain.

We define preparation as a process that: validates the consistency of the wasm binary (aka prevalidation) and the compilation of the wasm module into machine code (referred to as an artifact).

Besides pre-checking, preparation can also be triggered by execution, since a compiled artifact is needed for the execution. If an artifact already exists, execution will skip preparation. If it does do preparation, execution uses a more lenient timeout than preparation, to avoid the situation where honest validators fail on valid, pre-checked PVFs.

Runtime Architecture

It's clear that we want to separate different aspects of the runtime logic into different modules. Modules define their own storage, routines, and entry-points. They also define initialization and finalization logic.

Due to the (lack of) guarantees provided by a particular blockchain-runtime framework, there is no defined or dependable order in which modules' initialization or finalization logic will run. Supporting this blockchain-runtime framework is important enough to include that same uncertainty in our model of runtime modules in this guide. Furthermore, initialization logic of modules can trigger the entry-points or routines of other modules. This is one architectural pressure against dividing the runtime logic into multiple modules. However, in this case the benefits of splitting things up outweigh the costs, provided that we take certain precautions against initialization and entry-point races.

We also expect, although it's beyond the scope of this guide, that these runtime modules will exist alongside various other modules. This has two facets to consider. First, even if the modules that we describe here don't invoke each others' entry points or routines during initialization, we still have to protect against those other modules doing that. Second, some of those modules are expected to provide governance capabilities for the chain. Configuration exposed by parachain-host modules is mostly for the benefit of these governance modules, to allow the operators or community of the chain to tweak parameters.

The runtime's primary role is to manage scheduling and updating of parachains, as well as handling misbehavior reports and slashing. This guide doesn't focus on how parachains are registered, only that they are. Also, this runtime description assumes that validator sets are selected somehow, but doesn't assume any other details than a periodic session change event. Session changes give information about the incoming validator set and the validator set of the following session.

The runtime also serves another role, which is to make data available to the Node-side logic via Runtime APIs. These Runtime APIs should be sufficient for the Node-side code to author blocks correctly.

There is some functionality of the relay chain relating to parachains that we also consider beyond the scope of this document. In particular, all modules related to how parachains are registered aren't part of this guide, although we do provide routines that should be called by the registration process.

We will split the logic of the runtime up into these modules:

  • Initializer: manages initialization order of the other modules.
  • Shared: manages shared storage and configurations for other modules.
  • Configuration: manages configuration and configuration updates in a non-racy manner.
  • Paras: manages chain-head and validation code for parachains.
  • Scheduler: manages parachain scheduling as well as validator assignments.
  • Inclusion: handles the inclusion and availability of scheduled parachains.
  • SessionInfo: manages various session keys of validators and other params stored per session.
  • Disputes: handles dispute resolution for included, available parablocks.
  • Slashing: handles slashing logic for concluded disputes.
  • HRMP: handles horizontal messages between paras.
  • UMP: handles upward messages from a para to the relay chain.
  • DMP: handles downward messages from the relay chain to the para.

The Initializer module is special - it's responsible for handling the initialization logic of the other modules to ensure that the correct initialization order and related invariants are maintained. The other modules won't specify a on-initialize logic, but will instead expose a special semi-private routine that the initialization module will call. The other modules are relatively straightforward and perform the roles described above.

The Parachain Host operates under a changing set of validators. Time is split up into periodic sessions, where each session brings a potentially new set of validators. Sessions are buffered by one, meaning that the validators of the upcoming session n+1 are determined at the end of session n-1, right before session n starts. Parachain Host runtime modules need to react to changes in the validator set, as it will affect the runtime logic for processing candidate backing, availability bitfields, and misbehavior reports. The Parachain Host modules can't determine ahead-of-time exactly when session change notifications are going to happen within the block (note: this depends on module initialization order again - better to put session before parachains modules).

The relay chain is intended to use BABE or SASSAFRAS, which both have the property that a session changing at a block is determined not by the number of the block but instead by the time the block is authored. In some sense, sessions change in-between blocks, not at blocks. This has the side effect that the session of a child block cannot be determined solely by the parent block's identifier. Being able to unilaterally determine the validator-set at a specific block based on its parent hash would make a lot of Node-side logic much simpler.

In order to regain the property that the validator set of a block is predictable by its parent block, we delay session changes' application to Parachains by 1 block. This means that if there is a session change at block X, that session change will be stored and applied during initialization of direct descendants of X. This principal side effect of this change is that the Parachains runtime can disagree with session or consensus modules about which session it currently is. Misbehavior reporting routines in particular will be affected by this, although not severely. The parachains runtime might believe it is the last block of the session while the system is really in the first block of the next session. In such cases, a historical validator-set membership proof will need to accompany any misbehavior report, although they typically do not need to during current-session misbehavior reports.

So the other role of the initializer module is to forward session change notifications to modules in the initialization order. Session change is also the point at which the Configuration Module updates the configuration. Most of the other modules will handle changes in the configuration during their session change operation, so the initializer should provide both the old and new configuration to all the other modules alongside the session change notification. This means that a session change notification should consist of the following data:

#![allow(unused)]
fn main() {
struct SessionChangeNotification {
 // The new validators in the session.
 validators: Vec<ValidatorId>,
 // The validators for the next session.
 queued: Vec<ValidatorId>,
 // The configuration before handling the session change.
 prev_config: HostConfiguration,
 // The configuration after handling the session change.
 new_config: HostConfiguration,
 // A secure randomn seed for the session, gathered from BABE.
 random_seed: [u8; 32],
 // The session index of the beginning session.
 session_index: SessionIndex,
}
}

TODO Diagram: order of runtime operations (initialization, session change)

Initializer Pallet

This module is responsible for initializing the other modules in a deterministic order. It also has one other purpose as described in the overview of the runtime: accepting and forwarding session change notifications.

Storage

#![allow(unused)]
fn main() {
HasInitialized: bool;
// buffered session changes along with the block number at which they should be applied.
//
// typically this will be empty or one element long. ordered ascending by BlockNumber and insertion
// order.
BufferedSessionChanges: Vec<(BlockNumber, ValidatorSet, ValidatorSet)>;
}

Initialization

Before initializing modules, remove all changes from the BufferedSessionChanges with number less than or equal to the current block number, and apply the last one. The session change is applied to all modules in the same order as initialization.

The other parachains modules are initialized in this order:

  1. Configuration
  2. Shared
  3. Paras
  4. Scheduler
  5. Inclusion
  6. SessionInfo
  7. Disputes
  8. DMP
  9. UMP
  10. HRMP

The Configuration Module is first, since all other modules need to operate under the same configuration as each other. Then the Shared module is invoked, which determines the set of active validators. It would lead to inconsistency if, for example, the scheduler ran first and then the configuration was updated before the Inclusion module.

Set HasInitialized to true.

Session Change

Store the session change information in BufferedSessionChange along with the block number at which it was submitted, plus one. Although the expected operational parameters of the block authorship system should prevent more than one change from being buffered at any time, it may occur. Regardless, we always need to track the block number at which the session change can be applied so as to remain flexible over session change notifications being issued before or after initialization of the current block.

Finalization

Finalization order is less important in this case than initialization order, so we finalize the modules in the reverse order from initialization.

Set HasInitialized to false.

Configuration Pallet

This module is responsible for managing all configuration of the parachain host in-flight. It provides a central point for configuration updates to prevent races between configuration changes and parachain-processing logic. Configuration can only change during the session change routine, and as this module handles the session change notification first it provides an invariant that the configuration does not change throughout the entire session. Both the scheduler and inclusion modules rely on this invariant to ensure proper behavior of the scheduler.

The configuration that we will be tracking is the HostConfiguration struct.

Storage

The configuration module is responsible for two main pieces of storage.

#![allow(unused)]
fn main() {
/// The current configuration to be used.
Configuration: HostConfiguration;
/// A pending configuration to be applied on session change.
PendingConfigs: Vec<(SessionIndex, HostConfiguration)>;
/// A flag that says if the consistency checks should be omitted.
BypassConsistencyCheck: bool;
}

Session change

The session change routine works as follows:

  • If there is no pending configurations, then return early.
  • Take all pending configurations that are less than or equal to the current session index.
    • Get the pending configuration with the highest session index and apply it to the current configuration. Discard the earlier ones if any.

Routines

#![allow(unused)]
fn main() {
enum InconsistentError {
  // ...
}

impl HostConfiguration {
  fn check_consistency(&self) -> Result<(), InconsistentError> { /* ... */ }
}

/// Get the host configuration.
pub fn configuration() -> HostConfiguration {
  Configuration::get()
}

/// Schedules updating the host configuration. The update is given by the `updater` closure. The 
/// closure takes the current version of the configuration and returns the new version. 
/// Returns an `Err` if the closure returns a broken configuration. However, there are a couple of 
/// exceptions: 
///
/// - if the configuration that was passed in the closure is already broken, then it will pass the 
/// update: you cannot break something that is already broken.
/// - If the `BypassConsistencyCheck` flag is set, then the checks will be skipped.
///
/// The changes made by this function will always be scheduled at session X, where X is the current session index + 2.
/// If there is already a pending update for X, then the closure will receive the already pending configuration for 
/// session X.
///
/// If there is already a pending update for the current session index + 1, then it won't be touched. Otherwise,
/// that would violate the promise of this function that changes will be applied on the second session change (cur + 2).
fn schedule_config_update(updater: impl FnOnce(&mut HostConfiguration<BlockNumberFor<T>>)) -> DispatchResult
}

Entry-points

The Configuration module exposes an entry point for each configuration member. These entry-points accept calls only from governance origins. These entry-points will use the update_configuration routine to update the specific configuration field.

Shared Pallet

This module is responsible for managing shared storage and configuration for other modules.

It is important that other pallets are able to use the Shared Module, so it should not have a dependency on any other modules in the Parachains Runtime.

For the moment, it is used exclusively to track the current session index across the Parachains Runtime system, and when it should be allowed to schedule future changes to Paras or Configurations.

Constants

#![allow(unused)]
fn main() {
// `SESSION_DELAY` is used to delay any changes to Paras registration or configurations.
// Wait until the session index is 2 larger then the current index to apply any changes,
// which guarantees that at least one full session has passed before any changes are applied.
pub(crate) const SESSION_DELAY: SessionIndex = 2;
}

Storage

Helper structs:

#![allow(unused)]
fn main() {
struct AllowedRelayParentsTracker<Hash, BlockNumber> {
  // The past relay parents, paired with state roots, that are viable to build upon.
  //
  // They are in ascending chronologic order, so the newest relay parents are at
  // the back of the deque.
  //
  // (relay_parent, state_root)
  //
  // NOTE: the size limit of look-back is currently defined as a constant in Runtime.
  buffer: VecDeque<(Hash, Hash)>,

  // The number of the most recent relay-parent, if any.
  latest_number: BlockNumber,
}
}

Storage Layout:

#![allow(unused)]
fn main() {
/// The current session index within the Parachains Runtime system.
CurrentSessionIndex: SessionIndex;
/// All the validators actively participating in parachain consensus.
/// Indices are into the broader validator set.
ActiveValidatorIndices: Vec<ValidatorIndex>,
/// The parachain attestation keys of the validators actively participating in parachain consensus.
/// This should be the same length as `ActiveValidatorIndices`.
ActiveValidatorKeys: Vec<ValidatorId>
/// Relay-parents allowed to build candidates upon.
AllowedRelayParents: AllowedRelayParentsTracker<T::Hash, T::BlockNumber>,
}

Initialization

The Shared Module currently has no initialization routines.

The Shared Module is initialized directly after the Configuration module, but before all other modules. It is important to update the Shared Module before any other module since its state may be used within the logic of other modules, and it is important that the state is consistent across them.

Session Change

During a session change, the Shared Module receives and stores the current Session Index directly from the initializer module, along with the broader validator set, and it returns the new list of validators.

The list of validators should be first shuffled according to the chain's random seed and then truncated. The indices of these validators should be set to ActiveValidatorIndices and then returned back to the initializer. ActiveValidatorKeys should be set accordingly.

This information is used in the:

  • Configuration Module: For delaying updates to configurations until at lease one full session has passed.
  • Paras Module: For delaying updates to paras until at least one full session has passed.

Allowed relay parents buffer, which is maintained by ParaInherent module, is cleared on every session change.

Finalization

The Shared Module currently has no finalization routines.

Functions

  • scheduled_sessions() -> SessionIndex: Return the next session index where updates to the Parachains Runtime system would be safe to apply.
  • set_session_index(SessionIndex): For tests. Set the current session index in the Shared Module.

Disputes Pallet

After a backed candidate is made available, it is included and proceeds into an acceptance period during which validators are randomly selected to do (secondary) approval checks of the parablock. Any reports disputing the validity of the candidate will cause escalation, where even more validators are requested to check the block, and so on, until either the parablock is determined to be invalid or valid. Those on the wrong side of the dispute are slashed and, if the parablock is deemed invalid, the relay chain is rolled back to a point before that block was included.

However, this isn't the end of the story. We are working in a forkful blockchain environment, which carries three important considerations:

  1. For security, validators that misbehave shouldn't only be slashed on one fork, but on all possible forks. Validators that misbehave shouldn't be able to create a new fork of the chain when caught and get away with their misbehavior.
  2. It is possible (and likely) that the parablock being contested has not appeared on all forks.
  3. If a block author believes that there is a disputed parablock on a specific fork that will resolve to a reversion of the fork, that block author has more incentive to build on a different fork which does not include that parablock.

This means that in all likelihood, there is the possibility of disputes that are started on one fork of the relay chain, and as soon as the dispute resolution process starts to indicate that the parablock is indeed invalid, that fork of the relay chain will be abandoned and the dispute will never be fully resolved on that chain.

Even if this doesn't happen, there is the possibility that there are two disputes underway, and one resolves leading to a reversion of the chain before the other has concluded. In this case we want to both transplant the concluded dispute onto other forks of the chain as well as the unconcluded dispute.

We account for these requirements by having the disputes module handle two kinds of disputes.

  1. Local disputes: those contesting the validity of the current fork by disputing a parablock included within it.
  2. Remote disputes: a dispute that has partially or fully resolved on another fork which is transplanted to the local fork for completion and eventual slashing.

When a local dispute concludes negatively, the chain needs to be abandoned and reverted back to a block where the state does not contain the bad parablock. We expect that due to the Approval Checking Protocol, the current executing block should not be finalized. So we do two things when a local dispute concludes negatively:

  1. Freeze the state of parachains so nothing further is backed or included.
  2. Issue a digest in the header of the block that signals to nodes that this branch of the chain is to be abandoned.

If, as is expected, the chain is unfinalized, the freeze will have no effect as no honest validator will attempt to build on the frozen chain. However, if the approval checking protocol has failed and the bad parablock is finalized, the freeze serves to put the chain into a governance-only mode.

The storage of this module is designed around tracking DisputeStates, updating them with votes, and tracking blocks included by this branch of the relay chain. It also contains a Frozen parameter designed to freeze the state of all parachains.

Storage

Storage Layout:

#![allow(unused)]
fn main() {
LastPrunedSession: Option<SessionIndex>,

// All ongoing or concluded disputes for the last several sessions.
Disputes: double_map (SessionIndex, CandidateHash) -> Option<DisputeState>,
// All included blocks on the chain, as well as the block number in this chain that
// should be reverted back to if the candidate is disputed and determined to be invalid.
Included: double_map (SessionIndex, CandidateHash) -> Option<BlockNumber>,
// Whether the chain is frozen or not. Starts as `None`. When this is `Some`,
// the chain will not accept any new parachain blocks for backing or inclusion,
// and its value indicates the last valid block number in the chain.
// It can only be set back to `None` by governance intervention.
Frozen: Option<BlockNumber>,
}

byzantine_threshold refers to the maximum number f of validators which may be byzantine. The total number of validators is n = 3f + e where e in { 1, 2, 3 }.

Session Change

  1. If the current session is not greater than config.dispute_period + 1, nothing to do here.
  2. Set pruning_target = current_session - config.dispute_period - 1. We add the extra 1 because we want to keep things for config.dispute_period full sessions. The stuff at the end of the most recent session has been around for a little over 0 sessions, not a little over 1.
  3. If LastPrunedSession is None, then set LastPrunedSession to Some(pruning_target) and return.
  4. Otherwise, clear out all disputes and included candidates entries in the range last_pruned..=pruning_target and set LastPrunedSession to Some(pruning_target).

Block Initialization

This is currently a no op.

Routines

  • filter_multi_dispute_data(MultiDisputeStatementSet) -> MultiDisputeStatementSet:

    1. Takes a MultiDisputeStatementSet and filters it down to a MultiDisputeStatementSet that satisfies all the criteria of provide_multi_dispute_data. That is, eliminating ancient votes, duplicates and unconfirmed disputes. This can be used by block authors to create the final submission in a block which is guaranteed to pass the provide_multi_dispute_data checks.
  • provide_multi_dispute_data(MultiDisputeStatementSet) -> Vec<(SessionIndex, Hash)>:

    1. Pass on each dispute statement set to provide_dispute_data, propagating failure.
    2. Return a list of all candidates who just had disputes initiated.
  • provide_dispute_data(DisputeStatementSet) -> bool: Provide data to an ongoing dispute or initiate a dispute.

    1. All statements must be issued under the correct session for the correct candidate.
    2. SessionInfo is used to check statement signatures and this function should fail if any signatures are invalid.
    3. If there is no dispute under Disputes, create a new DisputeState with blank bitfields.
    4. If concluded_at is Some, and is concluded_at + config.post_conclusion_acceptance_period < now, return false.
    5. Import all statements into the dispute. This should fail if any statements are duplicate or if the corresponding bit for the corresponding validator is set in the dispute already.
    6. If concluded_at is None, reward all statements.
    7. If concluded_at is Some, reward all statements slightly less.
    8. If either side now has supermajority and did not previously, slash the other side. This may be both sides, and we support this possibility in code, but note that this requires validators to participate on both sides which has negative expected value. Set concluded_at to Some(now) if it was None.
    9. If just concluded against the candidate and the Included map contains (session, candidate): invoke revert_and_freeze with the stored block number.
    10. Return true if just initiated, false otherwise.
  • disputes() -> Vec<(SessionIndex, CandidateHash, DisputeState)>: Get a list of all disputes and info about dispute state.

    1. Iterate over all disputes in Disputes and collect into a vector.
  • note_included(SessionIndex, CandidateHash, included_in: BlockNumber):

    1. Add (SessionIndex, CandidateHash) to the Included map with included_in - 1 as the value.
    2. If there is a dispute under (SessionIndex, CandidateHash) that has concluded against the candidate, invoke revert_and_freeze with the stored block number.
  • concluded_invalid(SessionIndex, CandidateHash) -> bool: Returns whether a candidate has already concluded a dispute in the negative.

  • is_frozen(): Load the value of Frozen from storage. Return true if Some and false if None.

  • last_valid_block(): Load the value of Frozen from storage and return. None indicates that all blocks in the chain are potentially valid.

  • revert_and_freeze(BlockNumber):

    1. If is_frozen() return.
    2. Set Frozen to Some(BlockNumber) to indicate a rollback to the block number.
    3. Issue a Revert(BlockNumber + 1) log to indicate a rollback of the block's child in the header chain, which is the same as a rollback to the block number.

Disputes filtering

All disputes delivered to the runtime by the client are filtered before the actual import. In this context actual import means persisted in the runtime storage. The filtering has got two purposes:

  • Limit the amount of data saved onchain.
  • Prevent persisting malicious dispute data onchain.

Implementation note: Filtering is performed in function filter_dispute_data from Disputes pallet.

The filtering is performed on the whole statement set which is about to be imported onchain. The following filters are applied:

  1. Remove ancient disputes - if a dispute is concluded before the block number indicated in OLDEST_ACCEPTED parameter it is removed from the set. OLDEST_ACCEPTED is a runtime configuration option. Implementation note: dispute_post_conclusion_acceptance_period from HostConfiguration is used in the current Polkadot/Kusama implementation.
  2. Remove votes from unknown validators. If there is a vote from a validator which wasn't an authority in the session where the dispute was raised - they are removed. Please note that this step removes only single votes instead of removing the whole dispute.
  3. Remove one sided disputes - if a dispute doesn't contain two opposing votes it is not imported onchain. This serves as a measure not to import one sided disputes. A dispute is raised only if there are two opposing votes so if the client is not sending them the dispute is a potential spam.
  4. Remove unconfirmed disputes - if a dispute contains less votes than the byzantine threshold it is removed. This is also a spam precaution. A legitimate client will send only confirmed disputes to the runtime.

Rewards and slashing

After the disputes are filtered the validators participating in the disputes are rewarded and more importantly the offenders are slashed. Generally there can be two types of punishments:

  • "against valid" - the offender claimed that a valid candidate is invalid.
  • "for invalid" - the offender claimed that an invalid candidate is valid.

A dispute might be inconclusive. This means that it has timed out without being confirmed. A confirmed dispute is one containing votes more than the byzantine threshold (1/3 of the active validators). Validators participating in inconclusive disputes are not slashed. Thanks to the applied filtering (described in the previous section) one can be confident that there are no spam disputes in the runtime. So if a validator is not voting it is due to another reason (e.g. being under DoS attack). There is no reason to punish such validators with a slash.

Implementation note: Slashing is performed in process_checked_dispute_data from Disputes pallet.

Paras Pallet

The Paras module is responsible for storing information on parachains. Registered parachains cannot change except at session boundaries and after at least a full session has passed. This is primarily to ensure that the number and meaning of bits required for the availability bitfields does not change except at session boundaries.

It's also responsible for:

  • managing parachain validation code upgrades as well as maintaining availability of old parachain code and its pruning.
  • vetting PVFs by means of the PVF pre-checking mechanism.

Storage

Utility Structs

#![allow(unused)]
fn main() {
// the two key times necessary to track for every code replacement.
pub struct ReplacementTimes {
 /// The relay-chain block number that the code upgrade was expected to be activated.
 /// This is when the code change occurs from the para's perspective - after the
 /// first parablock included with a relay-parent with number >= this value.
 expected_at: BlockNumber,
 /// The relay-chain block number at which the parablock activating the code upgrade was
 /// actually included. This means considered included and available, so this is the time at which
 /// that parablock enters the acceptance period in this fork of the relay-chain.
 activated_at: BlockNumber,
}

/// Metadata used to track previous parachain validation code that we keep in
/// the state.
pub struct ParaPastCodeMeta {
 // Block numbers where the code was expected to be replaced and where the code
 // was actually replaced, respectively. The first is used to do accurate lookups
 // of historic code in historic contexts, whereas the second is used to do
 // pruning on an accurate timeframe. These can be used as indices
 // into the `PastCode` map along with the `ParaId` to fetch the code itself.
 upgrade_times: Vec<ReplacementTimes>,
 // This tracks the highest pruned code-replacement, if any.
 last_pruned: Option<BlockNumber>,
}

struct ParaGenesisArgs {
  /// The initial head-data to use.
  genesis_head: HeadData,
  /// The validation code to start with.
  validation_code: ValidationCode,
  /// True if parachain, false if parathread.
  parachain: bool,
}

/// The possible states of a para, to take into account delayed lifecycle changes.
pub enum ParaLifecycle {
  /// A Para is new and is onboarding.
  Onboarding,
  /// Para is a Parathread (on-demand parachain).
  Parathread,
  /// Para is a lease holding Parachain.
  Parachain,
  /// Para is a Parathread (on-demand Parachain) which is upgrading to a lease holding Parachain.
  UpgradingParathread,
  /// Para is a lease holding Parachain which is downgrading to an on-demand parachain.
  DowngradingParachain,
  /// Parathread (on-demand parachain) is being offboarded.
  OutgoingParathread,
  /// Parachain is being offboarded.
  OutgoingParachain,
}

enum PvfCheckCause {
  /// PVF vote was initiated by the initial onboarding process of the given para.
  Onboarding(ParaId),
  /// PVF vote was initiated by signalling of an upgrade by the given para.
  Upgrade {
    /// The ID of the parachain that initiated or is waiting for the conclusion of pre-checking.
    id: ParaId,
    /// The relay-chain block number that was used as the relay-parent for the parablock that
    /// initiated the upgrade.
    relay_parent_number: BlockNumber,
  },
}

struct PvfCheckActiveVoteState {
  // The two following vectors have their length equal to the number of validators in the active
  // set. They start with all zeroes. A 1 is set at an index when the validator at the that index
  // makes a vote. Once a 1 is set for either of the vectors, that validator cannot vote anymore.
  // Since the active validator set changes each session, the bit vectors are reinitialized as
  // well: zeroed and resized so that each validator gets its own bit.
  votes_accept: BitVec,
  votes_reject: BitVec,

  /// The number of session changes this PVF vote has observed. Therefore, this number is
  /// increased at each session boundary. When created, it is initialized with 0.
  age: SessionIndex,
  /// The block number at which this PVF vote was created.
  created_at: BlockNumber,
  /// A list of causes for this PVF pre-checking. Has at least one.
  causes: Vec<PvfCheckCause>,
}
}

Para Lifecycle

Because the state changes of parachains are delayed, we track the specific state of the para using the ParaLifecycle enum.

None         Parathread (on-demand parachain)    Parachain
 +                        +                          +
 |                        |                          |
 |   (≈2 Session Delay)   |                          |
 |                        |                          |
 +----------------------->+                          |
 |       Onboarding       |                          |
 |                        |                          |
 +-------------------------------------------------->+
 |       Onboarding       |                          |
 |                        |                          |
 |                        +------------------------->+
 |                        |   UpgradingParathread    |
 |                        |                          |
 |                        +<-------------------------+
 |                        |   DowngradingParachain   |
 |                        |                          |
 |<-----------------------+                          |
 |   OutgoingParathread   |                          |
 |                        |                          |
 +<--------------------------------------------------+
 |                        |    OutgoingParachain     |
 |                        |                          |
 +                        +                          +

Note that if PVF pre-checking is enabled, onboarding of a para may potentially be delayed. This can happen due to PVF pre-checking voting concluding late.

During the transition period, the para object is still considered in its existing state.

Storage Layout

#![allow(unused)]
fn main() {
use frame_system::pallet_prelude::BlockNumberFor;
/// All currently active PVF pre-checking votes.
///
/// Invariant:
/// - There are no PVF pre-checking votes that exists in list but not in the set and vice versa.
PvfActiveVoteMap: map ValidationCodeHash => PvfCheckActiveVoteState;
/// The list of all currently active PVF votes. Auxiliary to `PvfActiveVoteMap`.
PvfActiveVoteList: Vec<ValidationCodeHash>;
/// All parachains. Ordered ascending by ParaId. On-demand parachains are not included.
Parachains: Vec<ParaId>,
/// The current lifecycle state of all known Para Ids.
ParaLifecycle: map ParaId => Option<ParaLifecycle>,
/// The head-data of every registered para.
Heads: map ParaId => Option<HeadData>;
/// The context (relay-chain block number) of the most recent parachain head.
MostRecentContext: map ParaId => BlockNumber;
/// The validation code hash of every live para.
CurrentCodeHash: map ParaId => Option<ValidationCodeHash>;
/// Actual past code hash, indicated by the para id as well as the block number at which it became outdated.
PastCodeHash: map (ParaId, BlockNumber) => Option<ValidationCodeHash>;
/// Past code of parachains. The parachains themselves may not be registered anymore,
/// but we also keep their code on-chain for the same amount of time as outdated code
/// to keep it available for secondary checkers.
PastCodeMeta: map ParaId => ParaPastCodeMeta;
/// Which paras have past code that needs pruning and the relay-chain block at which the code was replaced.
/// Note that this is the actual height of the included block, not the expected height at which the
/// code upgrade would be applied, although they may be equal.
/// This is to ensure the entire acceptance period is covered, not an offset acceptance period starting
/// from the time at which the parachain perceives a code upgrade as having occurred.
/// Multiple entries for a single para are permitted. Ordered ascending by block number.
PastCodePruning: Vec<(ParaId, BlockNumber)>;
/// The block number at which the planned code change is expected for a para.
/// The change will be applied after the first parablock for this ID included which executes
/// in the context of a relay chain block with a number >= `expected_at`.
FutureCodeUpgrades: map ParaId => Option<BlockNumber>;
/// Hash of the actual future code of a para.
FutureCodeHash: map ParaId => Option<ValidationCodeHash>;
/// This is used by the relay-chain to communicate to a parachain a go-ahead with in the upgrade procedure.
///
/// This value is absent when there are no upgrades scheduled or during the time the relay chain
/// performs the checks. It is set at the first relay-chain block when the corresponding parachain
/// can switch its upgrade function. As soon as the parachain's block is included, the value
/// gets reset to `None`.
///
/// NOTE that this field is used by parachains via merkle storage proofs, therefore changing
/// the format will require migration of parachains.
UpgradeGoAheadSignal: map hasher(twox_64_concat) ParaId => Option<UpgradeGoAhead>;
/// This is used by the relay-chain to communicate that there are restrictions for performing
/// an upgrade for this parachain.
///
/// This may be a because the parachain waits for the upgrade cooldown to expire. Another
/// potential use case is when we want to perform some maintanance (such as storage migration)
/// we could restrict upgrades to make the process simpler.
///
/// NOTE that this field is used by parachains via merkle storage proofs, therefore changing
/// the format will require migration of parachains.
UpgradeRestrictionSignal: map hasher(twox_64_concat) ParaId => Option<UpgradeRestriction>;
/// The list of parachains that are awaiting for their upgrade restriction to cooldown.
///
/// Ordered ascending by block number.
UpgradeCooldowns: Vec<(ParaId, BlockNumberFor<T>)>;
/// The list of upcoming code upgrades. Each item is a pair of which para performs a code
/// upgrade and at which relay-chain block it is expected at.
///
/// Ordered ascending by block number.
UpcomingUpgrades: Vec<(ParaId, BlockNumberFor<T>)>;
/// The actions to perform during the start of a specific session index.
ActionsQueue: map SessionIndex => Vec<ParaId>;
/// Upcoming paras instantiation arguments.
///
/// NOTE that after PVF pre-checking is enabled the para genesis arg will have it's code set 
/// to empty. Instead, the code will be saved into the storage right away via `CodeByHash`.
UpcomingParasGenesis: map ParaId => Option<ParaGenesisArgs>;
/// The number of references on the validation code in `CodeByHash` storage.
CodeByHashRefs: map ValidationCodeHash => u32;
/// Validation code stored by its hash.
CodeByHash: map ValidationCodeHash => Option<ValidationCode>
}

Session Change

  1. Execute all queued actions for paralifecycle changes:
  2. Clean up outgoing paras.
    1. This means removing the entries under Heads, CurrentCode, FutureCodeUpgrades, FutureCode and MostRecentContext. An according entry should be added to PastCode, PastCodeMeta, and PastCodePruning using the outgoing ParaId and removed CurrentCode value. This is because any outdated validation code must remain available on-chain for a determined amount of blocks, and validation code outdated by de-registering the para is still subject to that invariant.
  3. Apply all incoming paras by initializing the Heads and CurrentCode using the genesis parameters as well as MostRecentContext to 0.
  4. Amend the Parachains list and ParaLifecycle to reflect changes in registered parachains.
  5. Amend the ParaLifecycle set to reflect changes in registered on-demand parachains.
  6. Upgrade all on-demand parachains that should become lease holding parachains, updating the Parachains list and ParaLifecycle.
  7. Downgrade all lease holding parachains that should become on-demand parachains, updating the Parachains list and ParaLifecycle.
  8. (Deferred) Return list of outgoing paras to the initializer for use by other modules.
  9. Go over all active PVF pre-checking votes:
  10. Increment age of the vote.
  11. If age reached cfg.pvf_voting_ttl, then enact PVF rejection and remove the vote from the active list.
  12. Otherwise, reinitialize the ballots.
    1. Resize the votes_accept/votes_reject to have the same length as the incoming validator set.
    2. Zero all the votes.

Initialization

  1. Do pruning based on all entries in PastCodePruning with BlockNumber <= now. Update the corresponding PastCodeMeta and PastCode accordingly.
  2. Toggle the upgrade related signals
  3. Collect all (para_id, expected_at) from UpcomingUpgrades where expected_at <= now and prune them. For each para pruned set UpgradeGoAheadSignal to GoAhead. Reserve weight for the state modification to upgrade each para pruned.
  4. Collect all (para_id, next_possible_upgrade_at) from UpgradeCooldowns where next_possible_upgrade_at <= now. For each para obtained this way reserve weight to remove its UpgradeRestrictionSignal on finalization.

Routines

  • schedule_para_initialize(ParaId, ParaGenesisArgs): Schedule a para to be initialized at the next session. Noop if para is already registered in the system with some ParaLifecycle.
  • schedule_para_cleanup(ParaId): Schedule a para to be cleaned up after the next full session.
  • schedule_parathread_upgrade(ParaId): Schedule a parathread (on-demand parachain) to be upgraded to a parachain.
  • schedule_parachain_downgrade(ParaId): Schedule a parachain to be downgraded from lease holding to on-demand.
  • schedule_code_upgrade(ParaId, new_code, relay_parent: BlockNumber, HostConfiguration): Schedule a future code upgrade of the given parachain. In case the PVF pre-checking is disabled, or the new code is already present in the storage, the upgrade will be applied after inclusion of a block of the same parachain executed in the context of a relay-chain block with number >= relay_parent + config.validation_upgrade_delay. If the upgrade is scheduled UpgradeRestrictionSignal is set and it will remain set until relay_parent + config.validation_upgrade_cooldown. In case the PVF pre-checking is enabled, or the new code is not already present in the storage, then the PVF pre-checking run will be scheduled for that validation code. If the pre-checking concludes with rejection, then the upgrade is canceled. Otherwise, after pre-checking is concluded the upgrade will be scheduled and be enacted as described above.
  • note_new_head(ParaId, HeadData, BlockNumber): note that a para has progressed to a new head, where the new head was executed in the context of a relay-chain block with given number, the latter value is inserted into the MostRecentContext mapping. This will apply pending code upgrades based on the block number provided. If an upgrade took place it will clear the UpgradeGoAheadSignal.
  • lifecycle(ParaId) -> Option<ParaLifecycle>: Return the ParaLifecycle of a para.
  • is_parachain(ParaId) -> bool: Returns true if the para ID references any live lease holding parachain, including those which may be transitioning to an on-demand parachain in the future.
  • is_parathread(ParaId) -> bool: Returns true if the para ID references any live parathread (on-demand parachain), including those which may be transitioning to a lease holding parachain in the future.
  • is_valid_para(ParaId) -> bool: Returns true if the para ID references either a live on-demand parachain or live lease holding parachain.
  • can_upgrade_validation_code(ParaId) -> bool: Returns true if the given para can signal code upgrade right now.
  • pvfs_require_prechecking() -> Vec<ValidationCodeHash>: Returns the list of PVF validation code hashes that require PVF pre-checking votes.

Finalization

Collect all (para_id, next_possible_upgrade_at) from UpgradeCooldowns where next_possible_upgrade_at <= now and prune them. For each para pruned remove its UpgradeRestrictionSignal.

Scheduler Pallet

TODO: this section is still heavily under construction. key questions about availability cores and validator assignment are still open and the flow of the the section may be contradictory or inconsistent

The Scheduler module is responsible for two main tasks:

  • Partitioning validators into groups and assigning groups to parachains.
  • Scheduling parachains for each block

It aims to achieve these tasks with these goals in mind:

  • It should be possible to know at least a block ahead-of-time, ideally more, which validators are going to be assigned to which parachains.
  • Parachains that have a candidate pending availability in this fork of the chain should not be assigned.
  • Validator assignments should not be gameable. Malicious cartels should not be able to manipulate the scheduler to assign themselves as desired.
  • High or close to optimal throughput of parachains. Work among validator groups should be balanced.

Availability Cores

The Scheduler manages resource allocation using the concept of "Availability Cores". There will be one availability core for each lease holding parachain, and a fixed number of cores used for multiplexing on-demand parachains. Validators will be partitioned into groups, with the same number of groups as availability cores. Validator groups will be assigned to different availability cores over time.

An availability core can exist in either one of two states at the beginning or end of a block: free or occupied. A free availability core can have a lease holding or on-demand parachain assigned to it for the potential to have a backed candidate included. After backing, the core enters the occupied state as the backed candidate is pending availability. There is an important distinction: a core is not considered occupied until it is in charge of a block pending availability, although the implementation may treat scheduled cores the same as occupied ones for brevity. A core exits the occupied state when the candidate is no longer pending availability - either on timeout or on availability. A core starting in the occupied state can move to the free state and back to occupied all within a single block, as availability bitfields are processed before backed candidates. At the end of the block, there is a possible timeout on availability which can move the core back to the free state if occupied.

Cores are treated as an ordered list and are typically referred to by their index in that list.

%3Availability Core State Machinevg1Freevg2Occupiedvg1->vg2Assignment & Backingvg2->vg1Availability or Timeout
%3Availability Core Transitions within Blockcluster_leftcluster_rightfr1Freefr2Freefr1->fr2No BackingoccOccupiedfr1->occBackingocc2Occupiedfr3Freeocc2->fr3Availabilityocc3Occupiedocc2->occ3No availabilityfr4Freefr3->fr4No backingocc4Occupiedfr3->occ4Backingocc3->fr3Availability Timeoutocc3->occ4(no change)

Validator Groups

Validator group assignments do not need to change very quickly. The security benefits of fast rotation are redundant with the challenge mechanism in the Approval process. Because of this, we only divide validators into groups at the beginning of the session and do not shuffle membership during the session. However, we do take steps to ensure that no particular validator group has dominance over a single lease holding parachain or on-demand parachain-multiplexer for an entire session to provide better guarantees of live-ness.

Validator groups rotate across availability cores in a round-robin fashion, with rotation occurring at fixed intervals. The i'th group will be assigned to the (i+k)%n'th core at any point in time, where k is the number of rotations that have occurred in the session, and n is the number of cores. This makes upcoming rotations within the same session predictable.

When a rotation occurs, validator groups are still responsible for distributing availability chunks for any previous cores that are still occupied and pending availability. In practice, rotation and availability-timeout frequencies should be set so this will only be the core they have just been rotated from. It is possible that a validator group is rotated onto a core which is currently occupied. In this case, the validator group will have nothing to do until the previously-assigned group finishes their availability work and frees the core or the availability process times out. Depending on if the core is for a lease holding parachain or on-demand parachain, a different timeout t from the HostConfiguration will apply. Availability timeouts should only be triggered in the first t-1 blocks after the beginning of a rotation.

Claims

On-demand parachains operate on a system of claims. Collators purchase claims on authoring the next block of an on-demand parachain, although the purchase mechanism is beyond the scope of the scheduler. The scheduler guarantees that they'll be given at least a certain number of attempts to author a candidate that is backed. Attempts that fail during the availability phase are not counted, since ensuring availability at that stage is the responsibility of the backing validators, not of the collator. When a claim is accepted, it is placed into a queue of claims, and each claim is assigned to a particular on-demand parachain-multiplexing core in advance. Given that the current assignments of validator groups to cores are known, and the upcoming assignments are predictable, it is possible for on-demand parachain collators to know who they should be talking to now and how they should begin establishing connections with as a fallback.

With this information, the Node-side can be aware of which on-demand parachains have a good chance of being includable within the relay-chain block and can focus any additional resources on backing candidates from those on-demand parachains. Furthermore, Node-side code is aware of which validator group will be responsible for that thread. If the necessary conditions are reached for core reassignment, those candidates can be backed within the same block as the core being freed.

On-demand claims, when scheduled onto a free core, may not result in a block pending availability. This may be due to collator error, networking timeout, or censorship by the validator group. In this case, the claims should be retried a certain number of times to give the collator a fair shot.

Storage

Utility structs:

#![allow(unused)]
fn main() {
// A claim on authoring the next block for a given parathread (on-demand parachain).
struct ParathreadClaim(ParaId, CollatorId);

// An entry tracking a parathread (on-demand parachain) claim to ensure it does not 
// pass the maximum number of retries.
struct ParathreadEntry {
  claim: ParathreadClaim,
  retries: u32,
}

// A queued parathread (on-demand parachain) entry, pre-assigned to a core.
struct QueuedParathread {
  claim: ParathreadEntry,
  /// offset within the set of parathreads (on-demand parachains) ranged `0..config.parathread_cores`.
  core_offset: u32,
}

struct ParathreadQueue {
  queue: Vec<QueuedParathread>,
  /// offset within the set of parathreads (on-demand parachains) ranged `0..config.parathread_cores`.
  next_core_offset: u32,
}

enum CoreOccupied {
  // On-demand parachain
  Parathread(ParathreadEntry), // claim & retries
  Parachain,
}

enum AssignmentKind {
  Parachain,
  // On-demand parachain
  Parathread(CollatorId, u32),
}

struct CoreAssignment {
  core: CoreIndex,
  para_id: ParaId,
  kind: AssignmentKind,
}
// reasons a core might be freed.
enum FreedReason {
  Concluded,
  TimedOut,
}
}

Storage layout:

#![allow(unused)]
fn main() {
/// All the validator groups. One for each core. Indices are into the `ActiveValidators` storage.
ValidatorGroups: Vec<Vec<ValidatorIndex>>;
/// A queue of upcoming parathread (on-demand parachain) claims and which core they should be mapped onto.
ParathreadQueue: ParathreadQueue;
/// One entry for each availability core. Entries are `None` if the core is not currently occupied.
/// The i'th parachain lease belongs to the i'th core, with the remaining cores all being
/// on-demand parachain-multiplexers.
AvailabilityCores: Vec<Option<CoreOccupied>>;
/// An index used to ensure that only one claim on a parathread (on-demand parachain) exists in the queue or is
/// currently being handled by an occupied core.
ParathreadClaimIndex: Vec<ParaId>;
/// The block number where the session start occurred. Used to track how many group rotations have occurred.
SessionStartBlock: BlockNumber;
/// Currently scheduled cores - free but up to be occupied.
/// The value contained here will not be valid after the end of a block. 
/// Runtime APIs should be used to determine scheduled cores
/// for the upcoming block.
Scheduled: Vec<CoreAssignment>, // sorted ascending by CoreIndex.
}

Session Change

Session changes are the only time that configuration can change, and the Configuration module's session-change logic is handled before this module's. We also lean on the behavior of the Inclusion module which clears all its occupied cores on session change. Thus we don't have to worry about cores being occupied across session boundaries and it is safe to re-size the AvailabilityCores bitfield.

Actions:

  1. Set SessionStartBlock to current block number + 1, as session changes are applied at the end of the block.
  2. Clear all Some members of AvailabilityCores. Return all parathread claims to queue with retries un-incremented.
  3. Set configuration = Configuration::configuration() (see HostConfiguration)
  4. Fetch Shared::ActiveValidators as AV.
  5. Determine the number of cores & validator groups as n_cores. This is the maximum of
    1. Paras::parachains().len() + configuration.parathread_cores
    2. n_validators / max_validators_per_core if configuration.max_validators_per_core is Some and non-zero.
  6. Resize AvailabilityCores to have length n_cores with all None entries.
  7. Compute new validator groups by shuffling using a secure randomness beacon
    • Note that the total number of validators V in AV may not be evenly divided by n_cores.
    • The groups are selected by partitioning AV. The first V % N groups will have (V / n_cores) + 1 members, while the remaining groups will have (V / N) members each.
    • Instead of using the indices within AV, which point to the broader set, indices into AV should be used. This implies that groups should have simply ascending validator indices.
  8. Prune the parathread (on-demand parachain) queue to remove all retries beyond configuration.parathread_retries.
    • Also prune all on-demand claims corresponding to de-registered parachains.
    • all pruned claims should have their entry removed from the parathread (on-demand parachain) index.
    • assign all non-pruned claims to new cores if the number of on-demand parachain cores has changed between the new_config and old_config of the SessionChangeNotification.
    • Assign claims in equal balance across all cores if rebalancing, and set the next_core of the ParathreadQueue (on-demand queue) by incrementing the relative index of the last assigned core and taking it modulo the number of on-demand cores.

Initialization

No initialization routine runs for this module.

Finalization

No finalization routine runs for this module.

Routines

  • add_parathread_claim(ParathreadClaim): Add a parathread (on-demand parachain) claim to the queue.
    • Fails if any on-demand claim on the same parachain is currently indexed.
    • Fails if the queue length is >= config.scheduling_lookahead * config.parathread_cores.
    • The core used for the on-demand claim is the next_core field of the ParathreadQueue (on-demand queue) and adding Paras::parachains().len() to it.
    • next_core is then updated by adding 1 and taking it modulo config.parathread_cores.
    • The claim is then added to the claim index.
  • free_cores(Vec<(CoreIndex, FreedReason)>): indicate previosuly-occupied cores which are to be considered returned and why they are being returned.
    • All freed lease holding parachain cores should be assigned to their respective parachain
    • All freed on-demand parachain cores whose reason for freeing was FreedReason::Concluded should have the claim removed from the claim index.
    • All freed on-demand cores whose reason for freeing was FreedReason::TimedOut should have the claim added to the parathread queue (on-demand queue) again without retries incremented
    • All freed on-demand cores should take the next on-demand parachain entry from the queue.
  • schedule(Vec<(CoreIndex, FreedReason)>, now: BlockNumber): schedule new core assignments, with a parameter indicating previously-occupied cores which are to be considered returned and why they are being returned.
    • Invoke free_cores(freed_cores)
    • The i'th validator group will be assigned to the (i+k)%n'th core at any point in time, where k is the number of rotations that have occurred in the session, and n is the total number of cores. This makes upcoming rotations within the same session predictable. Rotations are based off of now.
  • scheduled() -> Vec<CoreAssignment>: Get currently scheduled core assignments.
  • occupied(Vec<CoreIndex>). Note that the given cores have become occupied.
    • Behavior undefined if any given cores were not scheduled.
    • Behavior undefined if the given cores are not sorted ascending by core index
    • This clears them from Scheduled and marks each corresponding core in the AvailabilityCores as occupied.
    • Since both the availability cores and the newly-occupied cores lists are sorted ascending, this method can be implemented efficiently.
  • core_para(CoreIndex) -> ParaId: return the currently-scheduled or occupied ParaId for the given core.
  • group_validators(GroupIndex) -> Option<Vec<ValidatorIndex>>: return all validators in a given group, if the group index is valid for this session.
  • availability_timeout_predicate() -> Option<impl Fn(CoreIndex, BlockNumber) -> bool>: returns an optional predicate that should be used for timing out occupied cores. if None, no timing-out should be done. The predicate accepts the index of the core, and the block number since which it has been occupied. The predicate should be implemented based on the time since the last validator group rotation, and the respective parachain timeouts, i.e. only within max(config.chain_availability_period, config.thread_availability_period) of the last rotation would this return Some.
  • group_rotation_info(now: BlockNumber) -> GroupRotationInfo: Returns a helper for determining group rotation.
  • next_up_on_available(CoreIndex) -> Option<ScheduledCore>: Return the next thing that will be scheduled on this core assuming it is currently occupied and the candidate occupying it became available. Returns in ScheduledCore format (todo: link to Runtime APIs page; linkcheck doesn't allow this right now). For lease holding parachains, this is always the ID of the parachain and no specified collator. For on-demand parachains, this is based on the next item in the ParathreadQueue (on-demand queue) assigned to that core, and is None if there isn't one.
  • next_up_on_time_out(CoreIndex) -> Option<ScheduledCore>: Return the next thing that will be scheduled on this core assuming it is currently occupied and the candidate occupying it timed out. Returns in ScheduledCore format (todo: link to Runtime APIs page; linkcheck doesn't allow this right now). For parachains, this is always the ID of the parachain and no specified collator. For on-demand parachains, this is based on the next item in the ParathreadQueue (on-demand queue) assigned to that core, or if there isn't one, the claim that is currently occupying the core. Otherwise None.
  • clear():
    • Free all scheduled cores and return on-demand claims to queue, with retries incremented. Skip on-demand parachains which no longer exist under paras.

Inclusion Pallet

The inclusion module is responsible for inclusion and availability of scheduled parachains. It also manages the UMP dispatch queue of each parachain.

Storage

Helper structs:

#![allow(unused)]
fn main() {
struct AvailabilityBitfield {
  bitfield: BitVec, // one bit per core.
  submitted_at: BlockNumber, // for accounting, as meaning of bits may change over time.
}

struct CandidatePendingAvailability {
  core: CoreIndex, // availability core
  hash: CandidateHash,
  descriptor: CandidateDescriptor,
  availability_votes: Bitfield, // one bit per validator.
  relay_parent_number: BlockNumber, // number of the relay-parent.
  backers: Bitfield, // one bit per validator, set for those who backed the candidate.
  backed_in_number: BlockNumber,
  backing_group: GroupIndex,
}
}

Storage Layout:

#![allow(unused)]
fn main() {
/// The latest bitfield for each validator, referred to by index.
bitfields: map ValidatorIndex => AvailabilityBitfield;
/// Candidates pending availability.
PendingAvailability: map ParaId => CandidatePendingAvailability;
/// The commitments of candidates pending availability, by ParaId.
PendingAvailabilityCommitments: map ParaId => CandidateCommitments;
}

Config Dependencies

  • MessageQueue: The message queue provides general queueing and processing functionality. Currently it replaces the old UMP dispatch queue. Other use-cases can be implemented as well by adding new variants to AggregateMessageOrigin. Normally it should be set to an instance of the MessageQueue pallet.

Session Change

  1. Clear out all candidates pending availability.
  2. Clear out all validator bitfields.

Optional:

  1. The UMP queue of all outgoing paras can be "swept". This would prevent the dispatch queue from automatically being serviced. It is a consideration for the chain and specific behaviour is not defined.

Initialization

No initialization routine runs for this module. However, the initialization of the MessageQueue pallet will attempt to process any pending UMP messages.

Routines

All failed checks should lead to an unrecoverable error making the block invalid.

  • process_bitfields(expected_bits, Bitfields, core_lookup: Fn(CoreIndex) -> Option<ParaId>):

    1. Call sanitize_bitfields<true> and use the sanitized signed_bitfields from now on.
    2. Call sanitize_backed_candidates<true> and use the sanitized backed_candidates from now on.
    3. Apply each bit of bitfield to the corresponding pending candidate, looking up on-demand parachain cores using the core_lookup. Disregard bitfields that have a 1 bit for any free cores.
    4. For each applied bit of each availability-bitfield, set the bit for the validator in the CandidatePendingAvailability's availability_votes bitfield. Track all candidates that now have >2/3 of bits set in their availability_votes. These candidates are now available and can be enacted.
    5. For all now-available candidates, invoke the enact_candidate routine with the candidate and relay-parent number.
    6. Return a list of (CoreIndex, CandidateHash) from freed cores consisting of the cores where candidates have become available.
  • sanitize_bitfields<T: crate::inclusion::Config>( unchecked_bitfields: UncheckedSignedAvailabilityBitfields, disputed_bitfield: DisputedBitfield, expected_bits: usize, parent_hash: T::Hash, session_index: SessionIndex, validators: &[ValidatorId], full_check: FullCheck, ):

    1. check that disputed_bitfield has the same number of bits as the expected_bits, iff not return early with an empty vec.
    2. each of the below checks is for each bitfield. If a check does not pass the bitfield will be skipped.
    3. check that there are no bits set that reference a disputed candidate.
    4. check that the number of bits is equal to expected_bits.
    5. check that the validator index is strictly increasing (and thus also unique).
    6. check that the validator bit index is not out of bounds.
    7. check the validators signature, iff full_check=FullCheck::Yes.
  • sanitize_backed_candidates<T: crate::inclusion::Config, F: FnMut(usize, &BackedCandidate<T::Hash>) -> bool>( mut backed_candidates: Vec<BackedCandidate<T::Hash>>, candidate_has_concluded_invalid_dispute: F, scheduled: &[CoreAssignment], )

    1. filter out any backed candidates that have concluded invalid.
    2. filters backed candidates whom's paraid was scheduled by means of the provided scheduled parameter.
    3. sorts remaining candidates with respect to the core index assigned to them.
  • process_candidates(allowed_relay_parents, BackedCandidates, scheduled: Vec<CoreAssignment>, group_validators: Fn(GroupIndex) -> Option<Vec<ValidatorIndex>>):

    For details on AllowedRelayParentsTracker see documentation for Shared module.

    1. check that each candidate corresponds to a scheduled core and that they are ordered in the same order the cores appear in assignments in scheduled.
    2. check that scheduled is sorted ascending by CoreIndex, without duplicates.
    3. check that the relay-parent from each candidate receipt is one of the allowed relay-parents.
    4. check that there is no candidate pending availability for any scheduled ParaId.
    5. check that each candidate's validation_data_hash corresponds to a PersistedValidationData computed from the state of the context block.
    6. If the core assignment includes a specific collator, ensure the backed candidate is issued by that collator.
    7. Ensure that any code upgrade scheduled by the candidate does not happen within config.validation_upgrade_cooldown of Paras::last_code_upgrade(para_id, true), if any, comparing against the value of Paras::FutureCodeUpgrades for the given para ID.
    8. Check the collator's signature on the candidate data.
    9. check the backing of the candidate using the signatures and the bitfields, comparing against the validators assigned to the groups, fetched with the group_validators lookup, while group indices are computed by Scheduler according to group rotation info.
    10. call check_upward_messages(config, para, commitments.upward_messages) to check that the upward messages are valid.
    11. call Dmp::check_processed_downward_messages(para, commitments.processed_downward_messages) to check that the DMQ is properly drained.
    12. call Hrmp::check_hrmp_watermark(para, commitments.hrmp_watermark) for each candidate to check rules of processing the HRMP watermark.
    13. using Hrmp::check_outbound_hrmp(sender, commitments.horizontal_messages) ensure that the each candidate sent a valid set of horizontal messages
    14. create an entry in the PendingAvailability map for each backed candidate with a blank availability_votes bitfield.
    15. create a corresponding entry in the PendingAvailabilityCommitments with the commitments.
    16. Return a Vec<CoreIndex> of all scheduled cores of the list of passed assignments that a candidate was successfully backed for, sorted ascending by CoreIndex.
  • enact_candidate(relay_parent_number: BlockNumber, CommittedCandidateReceipt):

    1. If the receipt contains a code upgrade, Call Paras::schedule_code_upgrade(para_id, code, relay_parent_number, config).

    TODO: Note that this is safe as long as we never enact candidates where the relay parent is across a session boundary. In that case, which we should be careful to avoid with contextual execution, the configuration might have changed and the para may de-sync from the host's understanding of it.

    1. Reward all backing validators of each candidate, contained within the backers field.
    2. call receive_upward_messages for each backed candidate, using the UpwardMessages from the CandidateCommitments.
    3. call Dmp::prune_dmq with the para id of the candidate and the candidate's processed_downward_messages.
    4. call Hrmp::prune_hrmp with the para id of the candiate and the candidate's hrmp_watermark.
    5. call Hrmp::queue_outbound_hrmp with the para id of the candidate and the list of horizontal messages taken from the commitment,
    6. Call Paras::note_new_head using the HeadData from the receipt and relay_parent_number.
  • collect_pending:

    #![allow(unused)]
    fn main() {
      fn collect_pending(f: impl Fn(CoreIndex, BlockNumber) -> bool) -> Vec<CoreIndex> {
        // sweep through all paras pending availability. if the predicate returns true, when given the core index and
        // the block number the candidate has been pending availability since, then clean up the corresponding storage for that candidate and the commitments.
        // return a vector of cleaned-up core IDs.
      }
    }
    
  • force_enact(ParaId): Forcibly enact the candidate with the given ID as though it had been deemed available by bitfields. Is a no-op if there is no candidate pending availability for this para-id. This should generally not be used but it is useful during execution of Runtime APIs, where the changes to the state are expected to be discarded directly after.

  • candidate_pending_availability(ParaId) -> Option<CommittedCandidateReceipt>: returns the CommittedCandidateReceipt pending availability for the para provided, if any.

  • pending_availability(ParaId) -> Option<CandidatePendingAvailability>: returns the metadata around the candidate pending availability for the para, if any.

  • collect_disputed(disputed: Vec<CandidateHash>) -> Vec<CoreIndex>: Sweeps through all paras pending availability. If the candidate hash is one of the disputed candidates, then clean up the corresponding storage for that candidate and the commitments. Return a vector of cleaned-up core IDs.

These functions were formerly part of the UMP pallet:

  • check_upward_messages(P: ParaId, Vec<UpwardMessage>):
    1. Checks that the parachain is not currently offboarding and error otherwise.
    2. Checks that there are at most config.max_upward_message_num_per_candidate messages to be enqueued.
    3. Checks that no message exceeds config.max_upward_message_size.
    4. Checks that the total resulting queue size would not exceed co.
    5. Verify that queuing up the messages could not result in exceeding the queue's footprint according to the config items config.max_upward_queue_count and config.max_upward_queue_size. The queue's current footprint is provided in well_known_keys in order to facilitate oraclisation on to the para.

Candidate Enactment:

  • receive_upward_messages(P: ParaId, Vec<UpwardMessage>):
    1. Process each upward message M in order:
      1. Place in the dispatch queue according to its para ID (or handle it immediately).

ParaInherent

This module is responsible for providing all data given to the runtime by the block author to the various parachains modules. The entry-point is mandatory, in that it must be invoked exactly once within every block, and it is also "inherent", in that it is provided with no origin by the block author. The data within it carries its own authentication; i.e. the data takes the form of signed statements by validators. Invalid data will be filtered and not applied.

This module does not have the same initialization/finalization concerns as the others, as it only requires that entry points be triggered after all modules have initialized and that finalization happens after entry points are triggered. Both of these are assumptions we have already made about the runtime's order of operations, so this module doesn't need to be initialized or finalized by the Initializer.

There are a couple of important notes to the operations in this inherent as they relate to disputes.

  1. We don't accept bitfields or backed candidates if in "governance-only" mode from having a local dispute conclude on this fork.
  2. When disputes are initiated, we remove the block from pending availability. This allows us to roll back chains to the block before blocks are included as opposed to backing. It's important to do this before processing bitfields.
  3. Inclusion::collect_disputed is kind of expensive so it's important to gate this on whether there are actually any new disputes. Which should be never.
  4. And we don't accept parablocks that have open disputes or disputes that have concluded against the candidate. It's important to import dispute statements before backing, but this is already the case as disputes are imported before processing bitfields.

Storage

#![allow(unused)]
fn main() {
/// Whether the para inherent was included or not.
Included: Option<()>,
}
#![allow(unused)]
fn main() {
/// Scraped on chain votes to be used in disputes off-chain.
OnChainVotes: Option<ScrapedOnChainVotes>,
}

Finalization

  1. Take (get and clear) the value of Included. If it is not Some, throw an unrecoverable error.

Entry Points

  • enter: This entry-point accepts one parameter: ParaInherentData.
  • create_inherent: This entry-point accepts one parameter: InherentData.

Both entry points share mostly the same code. create_inherent will meaningfully limit inherent data to adhere to the weight limit, in addition to sanitizing any inputs and filtering out invalid data. Conceptually it is part of the block production. The enter call on the other hand is part of block import and consumes/imports the data previously produced by create_inherent.

In practice both calls process inherent data and apply it to the state. Block production and block import should arrive at the same new state. Hence we re-use the same logic to ensure this is the case.

The only real difference between the two is, that on create_inherent we actually need the processed and filtered inherent data to build the block, while on enter the processed data should for one be identical to the incoming inherent data (assuming honest block producers) and second it is irrelevant, as we are not building a block but just processing it, so the processed inherent data is simply dropped.

This also means that the enter function keeps data around for no good reason. This seems acceptable though as the size of a block is rather limited. Nevertheless if we ever wanted to optimize this we can easily implement an inherent collector that has two implementations, where one clones and stores the data and the other just passes it on.

DMP Pallet

A module responsible for Downward Message Processing (DMP). See Messaging Overview for more details.

Storage

Storage layout required for implementation of DMP.

#![allow(unused)]
fn main() {
/// The downward messages addressed for a certain para.
DownwardMessageQueues: map ParaId => Vec<InboundDownwardMessage>;
/// A mapping that stores the downward message queue MQC head for each para.
///
/// Each link in this chain has a form:
/// `(prev_head, B, H(M))`, where
/// - `prev_head`: is the previous head hash or zero if none.
/// - `B`: is the relay-chain block number in which a message was appended.
/// - `H(M)`: is the hash of the message being appended.
DownwardMessageQueueHeads: map ParaId => Hash;
}

Initialization

No initialization routine runs for this module.

Routines

Candidate Acceptance Function:

  • check_processed_downward_messages(P: ParaId, relay_parent_number: BlockNumber, processed_downward_messages: u32):
    1. Checks that processed_downward_messages is at least 1 if DownwardMessageQueues for P is not empty at the given relay_parent_number.
    2. Checks that DownwardMessageQueues for P is at least processed_downward_messages long.

Candidate Enactment:

  • prune_dmq(P: ParaId, processed_downward_messages: u32):
    1. Remove the first processed_downward_messages from the DownwardMessageQueues of P.

Utility routines.

queue_downward_message(P: ParaId, M: DownwardMessage):

  1. Check if the size of M exceeds the config.max_downward_message_size. If so, return an error.
  2. Wrap M into InboundDownwardMessage using the current block number for sent_at.
  3. Obtain a new MQC link for the resulting InboundDownwardMessage and replace DownwardMessageQueueHeads for P with the resulting hash.
  4. Add the resulting InboundDownwardMessage into DownwardMessageQueues for P.

Session Change

  1. For each P in outgoing_paras (generated by Paras::on_new_session):
    1. Remove all DownwardMessageQueues of P.
    2. Remove DownwardMessageQueueHeads for P.

HRMP Pallet

A module responsible for Horizontally Relay-routed Message Passing (HRMP). See Messaging Overview for more details.

Storage

HRMP related structs:

#![allow(unused)]
fn main() {
/// A description of a request to open an HRMP channel.
struct HrmpOpenChannelRequest {
    /// Indicates if this request was confirmed by the recipient.
    confirmed: bool,
    /// The amount that the sender supplied at the time of creation of this request.
    sender_deposit: Balance,
    /// The maximum message size that could be put into the channel.
    max_message_size: u32,
    /// The maximum number of messages that can be pending in the channel at once.
    max_capacity: u32,
    /// The maximum total size of the messages that can be pending in the channel at once.
    max_total_size: u32,
}

/// A metadata of an HRMP channel.
struct HrmpChannel {
    /// The amount that the sender supplied as a deposit when opening this channel.
    sender_deposit: Balance,
    /// The amount that the recipient supplied as a deposit when accepting opening this channel.
    recipient_deposit: Balance,
    /// The maximum number of messages that can be pending in the channel at once.
    max_capacity: u32,
    /// The maximum total size of the messages that can be pending in the channel at once.
    max_total_size: u32,
    /// The maximum message size that could be put into the channel.
    max_message_size: u32,
    /// The current number of messages pending in the channel.
    /// Invariant: should be less or equal to `max_capacity`.
    msg_count: u32,
    /// The total size in bytes of all message payloads in the channel.
    /// Invariant: should be less or equal to `max_total_size`.
    total_size: u32,
    /// A head of the Message Queue Chain for this channel. Each link in this chain has a form:
    /// `(prev_head, B, H(M))`, where
    /// - `prev_head`: is the previous value of `mqc_head` or zero if none.
    /// - `B`: is the [relay-chain] block number in which a message was appended
    /// - `H(M)`: is the hash of the message being appended.
    /// This value is initialized to a special value that consists of all zeroes which indicates
    /// that no messages were previously added.
    mqc_head: Option<Hash>,
}
}

HRMP related storage layout

#![allow(unused)]
fn main() {
/// The set of pending HRMP open channel requests.
///
/// The set is accompanied by a list for iteration.
///
/// Invariant:
/// - There are no channels that exists in list but not in the set and vice versa.
HrmpOpenChannelRequests: map HrmpChannelId => Option<HrmpOpenChannelRequest>;
HrmpOpenChannelRequestsList: Vec<HrmpChannelId>;

/// This mapping tracks how many open channel requests are inititated by a given sender para.
/// Invariant: `HrmpOpenChannelRequests` should contain the same number of items that has `(X, _)`
/// as the number of `HrmpOpenChannelRequestCount` for `X`.
HrmpOpenChannelRequestCount: map ParaId => u32;
/// This mapping tracks how many open channel requests were accepted by a given recipient para.
/// Invariant: `HrmpOpenChannelRequests` should contain the same number of items `(_, X)` with
/// `confirmed` set to true, as the number of `HrmpAcceptedChannelRequestCount` for `X`.
HrmpAcceptedChannelRequestCount: map ParaId => u32;

/// A set of pending HRMP close channel requests that are going to be closed during the session change.
/// Used for checking if a given channel is registered for closure.
///
/// The set is accompanied by a list for iteration.
///
/// Invariant:
/// - There are no channels that exists in list but not in the set and vice versa.
HrmpCloseChannelRequests: map HrmpChannelId => Option<()>;
HrmpCloseChannelRequestsList: Vec<HrmpChannelId>;

/// The HRMP watermark associated with each para.
/// Invariant:
/// - each para `P` used here as a key should satisfy `Paras::is_valid_para(P)` within a session.
HrmpWatermarks: map ParaId => Option<BlockNumber>;
/// HRMP channel data associated with each para.
/// Invariant:
/// - each participant in the channel should satisfy `Paras::is_valid_para(P)` within a session.
HrmpChannels: map HrmpChannelId => Option<HrmpChannel>;
/// Ingress/egress indexes allow to find all the senders and receivers given the opposite
/// side. I.e.
///
/// (a) ingress index allows to find all the senders for a given recipient.
/// (b) egress index allows to find all the recipients for a given sender.
///
/// Invariants:
/// - for each ingress index entry for `P` each item `I` in the index should present in `HrmpChannels`
///   as `(I, P)`.
/// - for each egress index entry for `P` each item `E` in the index should present in `HrmpChannels`
///   as `(P, E)`.
/// - there should be no other dangling channels in `HrmpChannels`.
/// - the vectors are sorted.
HrmpIngressChannelsIndex: map ParaId => Vec<ParaId>;
HrmpEgressChannelsIndex: map ParaId => Vec<ParaId>;
/// Storage for the messages for each channel.
/// Invariant: cannot be non-empty if the corresponding channel in `HrmpChannels` is `None`.
HrmpChannelContents: map HrmpChannelId => Vec<InboundHrmpMessage>;
/// Maintains a mapping that can be used to answer the question:
/// What paras sent a message at the given block number for a given reciever.
/// Invariants:
/// - The inner `Vec<ParaId>` is never empty.
/// - The inner `Vec<ParaId>` cannot store two same `ParaId`.
/// - The outer vector is sorted ascending by block number and cannot store two items with the same
///   block number.
HrmpChannelDigests: map ParaId => Vec<(BlockNumber, Vec<ParaId>)>;
}

Initialization

No initialization routine runs for this module.

Routines

Candidate Acceptance Function:

  • check_hrmp_watermark(P: ParaId, new_hrmp_watermark):
    1. new_hrmp_watermark should be strictly greater than the value of HrmpWatermarks for P (if any).
    2. new_hrmp_watermark must not be greater than the context's block number.
    3. new_hrmp_watermark should be either
      1. equal to the context's block number
      2. or in HrmpChannelDigests for P an entry with the block number should exist
  • check_outbound_hrmp(sender: ParaId, Vec<OutboundHrmpMessage>):
    1. Checks that there are at most config.hrmp_max_message_num_per_candidate messages.
    2. Checks that horizontal messages are sorted by ascending recipient ParaId and there is no two horizontal messages have the same recipient.
    3. For each horizontal message M with the channel C identified by (sender, M.recipient) check:
      1. exists
      2. M's payload size doesn't exceed a preconfigured limit C.max_message_size
      3. M's payload size summed with the C.total_size doesn't exceed a preconfigured limit C.max_total_size.
      4. C.msg_count + 1 doesn't exceed a preconfigured limit C.max_capacity.

Candidate Enactment:

  • queue_outbound_hrmp(sender: ParaId, Vec<OutboundHrmpMessage>):
    1. For each horizontal message HM with the channel C identified by (sender, HM.recipient):
      1. Append HM into HrmpChannelContents that corresponds to C with sent_at equals to the current block number.
      2. Locate or create an entry in HrmpChannelDigests for HM.recipient and append sender into the entry's list.
      3. Increment C.msg_count
      4. Increment C.total_size by HM's payload size
      5. Append a new link to the MQC and save the new head in C.mqc_head. Note that the current block number as of enactment is used for the link.
  • prune_hrmp(recipient, new_hrmp_watermark):
    1. From HrmpChannelDigests for recipient remove all entries up to an entry with block number equal to new_hrmp_watermark.
    2. From the removed digests construct a set of paras that sent new messages within the interval between the old and new watermarks.
    3. For each channel C identified by (sender, recipient) for each sender coming from the set, prune messages up to the new_hrmp_watermark.
    4. For each pruned message M from channel C:
      1. Decrement C.msg_count
      2. Decrement C.total_size by M's payload size.
    5. Set HrmpWatermarks for P to be equal to new_hrmp_watermark

    NOTE: That collecting digests can be inefficient and the time it takes grows very fast. Thanks to the aggressive parameterization this shouldn't be a big of a deal. If that becomes a problem consider introducing an extra dictionary which says at what block the given sender sent a message to the recipient.

Entry-points

The following entry-points are meant to be used for HRMP channel management.

Those entry-points are meant to be called from a parachain. origin is defined as the ParaId of the parachain executed the message.

  • hrmp_init_open_channel(recipient, proposed_max_capacity, proposed_max_message_size):
    1. Check that the origin is not recipient.
    2. Check that proposed_max_capacity is less or equal to config.hrmp_channel_max_capacity and greater than zero.
    3. Check that proposed_max_message_size is less or equal to config.hrmp_channel_max_message_size and greater than zero.
    4. Check that recipient is a valid para.
    5. Check that there is no existing channel for (origin, recipient) in HrmpChannels.
    6. Check that there is no existing open channel request (origin, recipient) in HrmpOpenChannelRequests.
    7. Check that the sum of the number of already opened HRMP channels by the origin (the size of the set found HrmpEgressChannelsIndex for origin) and the number of open requests by the origin (the value from HrmpOpenChannelRequestCount for origin) doesn't exceed the limit of channels (config.hrmp_max_parachain_outbound_channels or config.hrmp_max_parathread_outbound_channels) minus 1.
    8. Check that origin's balance is more or equal to config.hrmp_sender_deposit
    9. Reserve the deposit for the origin according to config.hrmp_sender_deposit
    10. Increase HrmpOpenChannelRequestCount by 1 for origin.
    11. Append (origin, recipient) to HrmpOpenChannelRequestsList.
    12. Add a new entry to HrmpOpenChannelRequests for (origin, recipient)
      1. Set sender_deposit to config.hrmp_sender_deposit
      2. Set max_capacity to proposed_max_capacity
      3. Set max_message_size to proposed_max_message_size
      4. Set max_total_size to config.hrmp_channel_max_total_size
    13. Send a downward message to recipient notifying about an inbound HRMP channel request.
      • The DM is sent using queue_downward_message.
      • The DM is represented by the HrmpNewChannelOpenRequest XCM message.
        • sender is set to origin,
        • max_message_size is set to proposed_max_message_size,
        • max_capacity is set to proposed_max_capacity.
  • hrmp_accept_open_channel(sender):
    1. Check that there is an existing request between (sender, origin) in HrmpOpenChannelRequests
      1. Check that it is not confirmed.
    2. Check that the sum of the number of inbound HRMP channels opened to origin (the size of the set found in HrmpIngressChannelsIndex for origin) and the number of accepted open requests by the origin (the value from HrmpAcceptedChannelRequestCount for origin) doesn't exceed the limit of channels (config.hrmp_max_parachain_inbound_channels or config.hrmp_max_parathread_inbound_channels) minus 1.
    3. Check that origin's balance is more or equal to config.hrmp_recipient_deposit.
    4. Reserve the deposit for the origin according to config.hrmp_recipient_deposit
    5. For the request in HrmpOpenChannelRequests identified by (sender, P), set confirmed flag to true.
    6. Increase HrmpAcceptedChannelRequestCount by 1 for origin.
    7. Send a downward message to sender notifying that the channel request was accepted.
      • The DM is sent using queue_downward_message.
      • The DM is represented by the HrmpChannelAccepted XCM message.
        • recipient is set to origin.
  • hrmp_cancel_open_request(ch):
    1. Check that origin is either ch.sender or ch.recipient
    2. Check that the open channel request ch exists.
    3. Check that the open channel request for ch is not confirmed.
    4. Remove ch from HrmpOpenChannelRequests and HrmpOpenChannelRequestsList
    5. Decrement HrmpAcceptedChannelRequestCount for ch.recipient by 1.
    6. Unreserve the deposit of ch.sender.
  • hrmp_close_channel(ch):
    1. Check that origin is either ch.sender or ch.recipient
    2. Check that HrmpChannels for ch exists.
    3. Check that ch is not in the HrmpCloseChannelRequests set.
    4. If not already there, insert a new entry Some(()) to HrmpCloseChannelRequests for ch and append ch to HrmpCloseChannelRequestsList.
    5. Send a downward message to the opposite party notifying about the channel closing.
      • The DM is sent using queue_downward_message.
      • The DM is represented by the HrmpChannelClosing XCM message with:
        • initator is set to origin,
        • sender is set to ch.sender,
        • recipient is set to ch.recipient.
      • The opposite party is ch.sender if origin is ch.recipient and ch.recipient if origin is ch.sender.

Session Change

  1. For each P in outgoing_paras (generated by Paras::on_new_session):
    1. Remove all inbound channels of P, i.e. (_, P),
    2. Remove all outbound channels of P, i.e. (P, _),
    3. Remove HrmpOpenChannelRequestCount for P
    4. Remove HrmpAcceptedChannelRequestCount for P.
    5. Remove HrmpOpenChannelRequests and HrmpOpenChannelRequestsList for (P, _) and (_, P).
      1. For each removed channel request C:
        1. Unreserve the sender's deposit if the sender is not present in outgoing_paras
        2. Unreserve the recipient's deposit if C is confirmed and the recipient is not present in outgoing_paras
  2. For each channel designator D in HrmpOpenChannelRequestsList we query the request R from HrmpOpenChannelRequests:
    1. if R.confirmed = true,
      1. if both D.sender and D.recipient are not offboarded.
      2. create a new channel C between (D.sender, D.recipient).
        1. Initialize the C.sender_deposit with R.sender_deposit and C.recipient_deposit with the value found in the configuration config.hrmp_recipient_deposit.
        2. Insert sender into the set HrmpIngressChannelsIndex for the recipient.
        3. Insert recipient into the set HrmpEgressChannelsIndex for the sender.
      3. decrement HrmpOpenChannelRequestCount for D.sender by 1.
      4. decrement HrmpAcceptedChannelRequestCount for D.recipient by 1.
      5. remove R
      6. remove D
  3. For each HRMP channel designator D in HrmpCloseChannelRequestsList
    1. remove the channel identified by D, if exists.
    2. remove D from HrmpCloseChannelRequests.
    3. remove D from HrmpCloseChannelRequestsList

To remove a HRMP channel C identified with a tuple (sender, recipient):

  1. Return C.sender_deposit to the sender.
  2. Return C.recipient_deposit to the recipient.
  3. Remove C from HrmpChannels.
  4. Remove C from HrmpChannelContents.
  5. Remove recipient from the set HrmpEgressChannelsIndex for sender.
  6. Remove sender from the set HrmpIngressChannelsIndex for recipient.

Session Info

For disputes and approvals, we need access to information about validator sets from prior sessions. We also often want easy access to the same information about the current session's validator set. This module aggregates and stores this information in a rolling window while providing easy APIs for access.

Storage

Helper structs:

#![allow(unused)]
fn main() {
struct SessionInfo {
    /// Validators in canonical ordering.
    ///
    /// NOTE: There might be more authorities in the current session, than `validators` participating
    /// in parachain consensus. See
    /// [`max_validators`](https://github.com/paritytech/polkadot/blob/a52dca2be7840b23c19c153cf7e110b1e3e475f8/runtime/parachains/src/configuration.rs#L148).
    ///
    /// `SessionInfo::validators` will be limited to to `max_validators` when set.
    validators: Vec<ValidatorId>,
    /// Validators' authority discovery keys for the session in canonical ordering.
    ///
    /// NOTE: The first `validators.len()` entries will match the corresponding validators in
    /// `validators`, afterwards any remaining authorities can be found. This is any authorities not
    /// participating in parachain consensus - see
    /// [`max_validators`](https://github.com/paritytech/polkadot/blob/a52dca2be7840b23c19c153cf7e110b1e3e475f8/runtime/parachains/src/configuration.rs#L148)
    #[cfg_attr(feature = "std", ignore_malloc_size_of = "outside type")]
    discovery_keys: Vec<AuthorityDiscoveryId>,
    /// The assignment keys for validators.
    ///
    /// NOTE: There might be more authorities in the current session, than validators participating
    /// in parachain consensus. See
    /// [`max_validators`](https://github.com/paritytech/polkadot/blob/a52dca2be7840b23c19c153cf7e110b1e3e475f8/runtime/parachains/src/configuration.rs#L148).
    ///
    /// Therefore:
    /// ```ignore
    ///		assignment_keys.len() == validators.len() && validators.len() <= discovery_keys.len()
    ///	```
    assignment_keys: Vec<AssignmentId>,
    /// Validators in shuffled ordering - these are the validator groups as produced
    /// by the `Scheduler` module for the session and are typically referred to by
    /// `GroupIndex`.
    validator_groups: Vec<Vec<ValidatorIndex>>,
    /// The number of availability cores used by the protocol during this session.
    n_cores: u32,
    /// The zeroth delay tranche width.
    zeroth_delay_tranche_width: u32,
    /// The number of samples we do of `relay_vrf_modulo`.
    relay_vrf_modulo_samples: u32,
    /// The number of delay tranches in total.
    n_delay_tranches: u32,
    /// How many slots (BABE / SASSAFRAS) must pass before an assignment is considered a
    /// no-show.
    no_show_slots: u32,
    /// The number of validators needed to approve a block.
    needed_approvals: u32,
}
}

Storage Layout:

#![allow(unused)]
fn main() {
/// The earliest session for which previous session info is stored.
EarliestStoredSession: SessionIndex,
/// Session information. Should have an entry from `EarliestStoredSession..=CurrentSessionIndex`
Sessions: map SessionIndex => Option<SessionInfo>,
}

Session Change

  1. Update EarliestStoredSession based on config.dispute_period and remove all entries from Sessions from the previous value up to the new value.
  2. Create a new entry in Sessions with information about the current session. Use shared::ActiveValidators to determine the indices into the broader validator sets (validation, assignment, discovery) which are actually used for parachain validation. Only these validators should appear in the SessionInfo.

Routines

  • earliest_stored_session() -> SessionIndex: Yields the earliest session for which we have information stored.
  • session_info(session: SessionIndex) -> Option<SessionInfo>: Yields the session info for the given session, if stored.

Runtime APIs

Runtime APIs are the means by which the node-side code extracts information from the state of the runtime.

Every block in the relay-chain contains a state root which is the root hash of a state trie encapsulating all storage of runtime modules after execution of the block. This is a cryptographic commitment to a unique state. We use the terminology of accessing the state at a block to refer accessing the state referred to by the state root of that block.

Although Runtime APIs are often used for simple storage access, they are actually empowered to do arbitrary computation. The implementation of the Runtime APIs lives within the Runtime as Wasm code and exposes extern functions that can be invoked with arguments and have a return value. Runtime APIs have access to a variety of host functions, which are contextual functions provided by the Wasm execution context, that allow it to carry out many different types of behaviors.

Abilities provided by host functions includes:

  • State Access
  • Offchain-DB Access
  • Submitting transactions to the transaction queue
  • Optimized versions of cryptographic functions
  • More

So it is clear that Runtime APIs are a versatile and powerful tool to leverage the state of the chain. In general, we will use Runtime APIs for these purposes:

  • Access of a storage item
  • Access of a bundle of related storage items
  • Deriving a value from storage based on arguments
  • Submitting misbehavior reports

More broadly, we have the goal of using Runtime APIs to write Node-side code that fulfills the requirements set by the Runtime. In particular, the constraints set forth by the Scheduler and Inclusion modules. These modules are responsible for advancing paras with a two-phase protocol where validators are first chosen to validate and back a candidate and then required to ensure availability of referenced data. In the second phase, validators are meant to attest to those para-candidates that they have their availability chunk for. As the Node-side code needs to generate the inputs into these two phases, the runtime API needs to transmit information from the runtime that is aware of the Availability Cores model instantiated by the Scheduler and Inclusion modules.

Node-side code is also responsible for detecting and reporting misbehavior performed by other validators, and the set of Runtime APIs needs to provide methods for observing live disputes and submitting reports as transactions.

The next sections will contain information on specific runtime APIs. The format is this:

#![allow(unused)]
fn main() {
/// Fetch the value of the runtime API at the block.
///
/// Definitionally, the `at` parameter cannot be any block that is not in the chain.
/// Thus the return value is unconditional. However, for in-practice implementations
/// it may be possible to provide an `at` parameter as a hash, which may not refer to a
/// valid block or one which implements the runtime API. In those cases it would be
/// best for the implementation to return an error indicating the failure mode.
fn some_runtime_api(at: Block, arg1: Type1, arg2: Type2, ...) -> ReturnValue;
}

Certain runtime APIs concerning the state of a para require the caller to provide an OccupiedCoreAssumption. This indicates how the result of the runtime API should be computed if there is a candidate from the para occupying an availability core in the Inclusion Module.

The choices of assumption are whether the candidate occupying that core should be assumed to have been made available and included or timed out and discarded, along with a third option to assert that the core was not occupied. This choice affects everything from the parent head-data, the validation code, and the state of message-queues. Typically, users will take the assumption that either the core was free or that the occupying candidate was included, as timeouts are expected only in adversarial circumstances and even so, only in a small minority of blocks directly following validator set rotations.

#![allow(unused)]
fn main() {
/// An assumption being made about the state of an occupied core.
enum OccupiedCoreAssumption {
    /// The candidate occupying the core was made available and included to free the core.
    Included,
    /// The candidate occupying the core timed out and freed the core without advancing the para.
    TimedOut,
    /// The core was not occupied to begin with.
    Free,
}
}

Validators

Yields the validator-set at the state of a given block. This validator set is always the one responsible for backing parachains in the child of the provided block.

#![allow(unused)]
fn main() {
fn validators(at: Block) -> Vec<ValidatorId>;
}

Validator Groups

Yields the validator groups used during the current session. The validators in the groups are referred to by their index into the validator-set and this is assumed to be as-of the child of the block whose state is being queried.

#![allow(unused)]
fn main() {
/// A helper data-type for tracking validator-group rotations.
struct GroupRotationInfo {
    session_start_block: BlockNumber,
    group_rotation_frequency: BlockNumber,
    now: BlockNumber, // The successor of the block in whose state this runtime API is queried.
}

impl GroupRotationInfo {
    /// Returns the index of the group needed to validate the core at the given index,
    /// assuming the given amount of cores/groups.
    fn group_for_core(&self, core_index, cores) -> GroupIndex;

    /// Returns the block number of the next rotation after the current block. If the current block
    /// is 10 and the rotation frequency is 5, this should return 15.
    fn next_rotation_at(&self) -> BlockNumber;

    /// Returns the block number of the last rotation before or including the current block. If the
    /// current block is 10 and the rotation frequency is 5, this should return 10.
    fn last_rotation_at(&self) -> BlockNumber;
}

/// Returns the validator groups and rotation info localized based on the block whose state
/// this is invoked on. Note that `now` in the `GroupRotationInfo` should be the successor of
/// the number of the block.
fn validator_groups(at: Block) -> (Vec<Vec<ValidatorIndex>>, GroupRotationInfo);
}

Availability Cores

Yields information on all availability cores. Cores are either free or occupied. Free cores can have paras assigned to them. Occupied cores don't, but they can become available part-way through a block due to bitfields and then have something scheduled on them. To allow optimistic validation of candidates, the occupied cores are accompanied by information on what is upcoming. This information can be leveraged when validators perceive that there is a high likelihood of a core becoming available based on bitfields seen, and then optimistically validate something that would become scheduled based on that, although there is no guarantee on what the block producer will actually include in the block.

See also the Scheduler Module for a high-level description of what an availability core is and why it exists.

#![allow(unused)]
fn main() {
fn availability_cores(at: Block) -> Vec<CoreState>;
}

This is all the information that a validator needs about scheduling for the current block. It includes all information on Scheduler core-assignments and Inclusion state of blocks occupying availability cores. It includes data necessary to determine not only which paras are assigned now, but which cores are likely to become freed after processing bitfields, and exactly which bitfields would be necessary to make them so. The implementation of this runtime API should invoke Scheduler::clear and Scheduler::schedule(Vec::new(), current_block_number + 1) to ensure that scheduling is accurate.

#![allow(unused)]
fn main() {
struct OccupiedCore {
    // NOTE: this has no ParaId as it can be deduced from the candidate descriptor.
    /// If this core is freed by availability, this is the assignment that is next up on this
    /// core, if any. None if there is nothing queued for this core.
    next_up_on_available: Option<ScheduledCore>,
    /// The relay-chain block number this began occupying the core at.
    occupied_since: BlockNumber,
    /// The relay-chain block this will time-out at, if any.
    time_out_at: BlockNumber,
    /// If this core is freed by being timed-out, this is the assignment that is next up on this
    /// core. None if there is nothing queued for this core or there is no possibility of timing
    /// out.
    next_up_on_time_out: Option<ScheduledCore>,
    /// A bitfield with 1 bit for each validator in the set. `1` bits mean that the corresponding
    /// validators has attested to availability on-chain. A 2/3+ majority of `1` bits means that
    /// this will be available.
    availability: Bitfield,
    /// The group assigned to distribute availability pieces of this candidate.
    group_responsible: GroupIndex,
    /// The hash of the candidate occupying the core.
    candidate_hash: CandidateHash,
    /// The descriptor of the candidate occupying the core.
    candidate_descriptor: CandidateDescriptor,
}

struct ScheduledCore {
    /// The ID of a para scheduled.
    para_id: ParaId,
    /// The collator required to author the block, if any.
    collator: Option<CollatorId>,
}

enum CoreState {
    /// The core is currently occupied.
    Occupied(OccupiedCore),
    /// The core is currently free, with a para scheduled and given the opportunity
    /// to occupy.
    ///
    /// If a particular Collator is required to author this block, that is also present in this
    /// variant.
    Scheduled(ScheduledCore),
    /// The core is currently free and there is nothing scheduled. This can be the case for on-demand
    /// cores when there are no on-demand parachain blocks queued. Leased cores will never be left idle.
    Free,
}
}

Persisted Validation Data

Yields the PersistedValidationData for the given ParaId along with an assumption that should be used if the para currently occupies a core:

#![allow(unused)]
fn main() {
/// Returns the persisted validation data for the given para and occupied core assumption.
///
/// Returns `None` if either the para is not registered or the assumption is `Freed`
/// and the para already occupies a core.
fn persisted_validation_data(at: Block, ParaId, OccupiedCoreAssumption) -> Option<PersistedValidationData>;
}

Session Index

Get the session index that is expected at the child of a block.

In the Initializer module, session changes are buffered by one block. The session index of the child of any relay block is always predictable by that block's state.

This session index can be used to derive a SigningContext.

#![allow(unused)]
fn main() {
/// Returns the session index expected at a child of the block.
fn session_index_for_child(at: Block) -> SessionIndex;
}

Validation Code

Fetch the validation code used by a para, making the given OccupiedCoreAssumption.

#![allow(unused)]
fn main() {
fn validation_code(at: Block, ParaId, OccupiedCoreAssumption) -> Option<ValidationCode>;
}

Fetch the validation code (past, present or future) by its hash.

#![allow(unused)]
fn main() {
fn validation_code_by_hash(at: Block, ValidationCodeHash) -> Option<ValidationCode>;
}

Fetch the validation code hash used by a para, making the given OccupiedCoreAssumption.

⚠️ This API was introduced in ParachainHost v2.

#![allow(unused)]
fn main() {
fn validation_code_hash(at: Block, ParaId, OccupiedCoreAssumption) -> Option<ValidationCodeHash>;
}

Candidate Pending Availability

Get the receipt of a candidate pending availability. This returns Some for any paras assigned to occupied cores in availability_cores and None otherwise.

#![allow(unused)]
fn main() {
fn candidate_pending_availability(at: Block, ParaId) -> Option<CommittedCandidateReceipt>;
}

Candidate Events

Yields a vector of events concerning candidates that occurred within the given block.

#![allow(unused)]
fn main() {
enum CandidateEvent {
	/// This candidate receipt was backed in the most recent block.
	CandidateBacked(CandidateReceipt, HeadData, CoreIndex, GroupIndex),
	/// This candidate receipt was included and became a parablock at the most recent block.
	CandidateIncluded(CandidateReceipt, HeadData, CoreIndex, GroupIndex),
	/// This candidate receipt was not made available in time and timed out.
	CandidateTimedOut(CandidateReceipt, HeadData, CoreIndex),
}

fn candidate_events(at: Block) -> Vec<CandidateEvent>;
}

Disputes Info

Get information about all disputes known by the chain as well as information about which validators the disputes subsystem will accept disputes from. These disputes may be either live or concluded. The DisputeState can be used to determine whether the dispute still accepts votes, as well as which validators' votes may be included.

#![allow(unused)]
fn main() {
struct Dispute {
    session: SessionIndex,
    candidate: CandidateHash,
    dispute_state: DisputeState,
    local: bool,
}

struct SpamSlotsInfo {
    max_spam_slots: u32,
    session_spam_slots: Vec<(SessionIndex, Vec<u32>)>,
}

struct DisputesInfo {
    disputes: Vec<Dispute>,
    spam_slots: SpamSlotsInfo,
}

fn disputes_info() -> DisputesInfo;
}

Candidates Included

This runtime API is for checking which candidates have been included within the chain, locally.

#![allow(unused)]
fn main() {
/// Input and output have the same length.
fn candidates_included(Vec<(SessionIndex, CandidateHash)>) -> Vec<bool>;
}

PVF Pre-checking

⚠️ This runtime API was added in v2.

There are two main runtime APIs to work with PVF pre-checking.

The first runtime API is designed to fetch all PVFs that require pre-checking voting. The PVFs are identified by their code hashes. As soon as the PVF gains required support, the runtime API will not return the PVF anymore.

#![allow(unused)]
fn main() {
fn pvfs_require_precheck() -> Vec<ValidationCodeHash>;
}

The second runtime API is needed to submit the judgement for a PVF, whether it is approved or not. The voting process uses unsigned transactions. The PvfCheckStatement is circulated through the network via gossip similar to a normal transaction. At some point the validator will include the statement in the block, where it will be processed by the runtime. If that was the last vote before gaining the super-majority, this PVF will not be returned by pvfs_require_precheck anymore.

#![allow(unused)]
fn main() {
fn submit_pvf_check_statement(stmt: PvfCheckStatement, signature: ValidatorSignature);
}

Node Architecture

Design Goals

  • Modularity: Components of the system should be as self-contained as possible. Communication boundaries between components should be well-defined and mockable. This is key to creating testable, easily reviewable code.
  • Minimizing side effects: Components of the system should aim to minimize side effects and to communicate with other components via message-passing.
  • Operational Safety: The software will be managing signing keys where conflicting messages can lead to large amounts of value to be slashed. Care should be taken to ensure that no messages are signed incorrectly or in conflict with each other.

The architecture of the node-side behavior aims to embody the Rust principles of ownership and message-passing to create clean, isolatable code. Each resource should have a single owner, with minimal sharing where unavoidable.

Many operations that need to be carried out involve the network, which is asynchronous. This asynchrony affects all core subsystems that rely on the network as well. The approach of hierarchical state machines is well-suited to this kind of environment.

We introduce

Components

The node architecture consists of the following components:

  • The Overseer (and subsystems): A hierarchy of state machines where an overseer supervises subsystems. Subsystems can contain their own internal hierarchy of jobs. This is elaborated on in the next section on Subsystems.
  • A block proposer: Logic triggered by the consensus algorithm of the chain when the node should author a block.
  • A GRANDPA voting rule: A strategy for selecting chains to vote on in the GRANDPA algorithm to ensure that only valid parachain candidates appear in finalized relay-chain blocks.

Assumptions

The Node-side code comes with a set of assumptions that we build upon. These assumptions encompass most of the fundamental blockchain functionality.

We assume the following constraints regarding provided basic functionality:

  • The underlying consensus algorithm, whether it is BABE or SASSAFRAS is implemented.
  • There is a chain synchronization protocol which will search for and download the longest available chains at all times.
  • The state of all blocks at the head of the chain is available. There may be state pruning such that state of the last k blocks behind the last finalized block are available, as well as the state of all their descendants. This assumption implies that the state of all active leaves and their last k ancestors are all available. The underlying implementation is expected to support k of a few hundred blocks, but we reduce this to a very conservative k=5 for our purposes.
  • There is an underlying networking framework which provides peer discovery services which will provide us with peers and will not create "loopback" connections to our own node. The number of peers we will have is assumed to be bounded at 1000.
  • There is a transaction pool and a transaction propagation mechanism which maintains a set of current transactions and distributes to connected peers. Current transactions are those which are not outdated relative to some "best" fork of the chain, which is part of the active heads, and have not been included in the best fork.

Subsystems and Jobs

In this section we define the notions of Subsystems and Jobs. These are guidelines for how we will employ an architecture of hierarchical state machines. We'll have a top-level state machine which oversees the next level of state machines which oversee another layer of state machines and so on. The next sections will lay out these guidelines for what we've called subsystems and jobs, since this model applies to many of the tasks that the Node-side behavior needs to encompass, but these are only guidelines and some Subsystems may have deeper hierarchies internally.

Subsystems are long-lived worker tasks that are in charge of performing some particular kind of work. All subsystems can communicate with each other via a well-defined protocol. Subsystems can't generally communicate directly, but must coordinate communication through an Overseer, which is responsible for relaying messages, handling subsystem failures, and dispatching work signals.

Most work that happens on the Node-side is related to building on top of a specific relay-chain block, which is contextually known as the "relay parent". We call it the relay parent to explicitly denote that it is a block in the relay chain and not on a parachain. We refer to the parent because when we are in the process of building a new block, we don't know what that new block is going to be. The parent block is our only stable point of reference, even though it is usually only useful when it is not yet a parent but in fact a leaf of the block-DAG expected to soon become a parent (because validators are authoring on top of it). Furthermore, we are assuming a forkful blockchain-extension protocol, which means that there may be multiple possible children of the relay-parent. Even if the relay parent has multiple children blocks, the parent of those children is the same, and the context in which those children is authored should be the same. The parent block is the best and most stable reference to use for defining the scope of work items and messages, and is typically referred to by its cryptographic hash.

Since this goal of determining when to start and conclude work relative to a specific relay-parent is common to most, if not all subsystems, it is logically the job of the Overseer to distribute those signals as opposed to each subsystem duplicating that effort, potentially being out of synchronization with each other. Subsystem A should be able to expect that subsystem B is working on the same relay-parents as it is. One of the Overseer's tasks is to provide this heartbeat, or synchronized rhythm, to the system.

The work that subsystems spawn to be done on a specific relay-parent is known as a job. Subsystems should set up and tear down jobs according to the signals received from the overseer. Subsystems may share or cache state between jobs.

Subsystems must be robust to spurious exits. The outputs of the set of subsystems as a whole comprises of signed messages and data committed to disk. Care must be taken to avoid issuing messages that are not substantiated. Since subsystems need to be safe under spurious exits, it is the expected behavior that an OverseerSignal::Conclude can just lead to breaking the loop and exiting directly as opposed to waiting for everything to shut down gracefully.

Subsystem Message Traffic

Which subsystems send messages to which other subsystems.

Note: This diagram omits the overseer for simplicity. In fact, all messages are relayed via the overseer.

Note: Messages with a filled diamond arrowhead ("♦") include a oneshot::Sender which communicates a response from the recipient. Messages with an open triangle arrowhead ("Δ") do not include a return sender.

%3av_storeAvailability Storechn_apiChain APIav_store->chn_apiBlockNumberav_store->chn_apiBlockHeaderav_store->chn_apiFinalizedBlockHashrunt_apiRuntime APIav_store->runt_apiRequest::CandidateEventsav_store->runt_apiRequest::Validatorsavail_distAvailability Distributionavail_dist->av_storeQueryDataAvailabilityavail_dist->av_storeQueryChunkavail_dist->av_storeStoreChunkavail_dist->chn_apiAncestorsnet_brdgNetwork Bridgeavail_dist->net_brdgRequest::SendValidationMessagesavail_dist->net_brdgReportPeeravail_dist->runt_apiRequest::AvailabilityCoresavail_dist->runt_apiRequest::Validatorsavail_dist->runt_apiRequest::SessionIndexForChildavail_rcovAvailability Recoveryavail_rcov->av_storeQueryChunkavail_rcov->net_brdgReportPeeravail_rcov->net_brdgConnectToValidatorsavail_rcov->net_brdgSendValidationMessage::Chunkavail_rcov->net_brdgSendValidationMessage::RequestChunkbitf_distBitfield Distributionbitf_dist->net_brdgReportPeerbitf_dist->net_brdgSendValidationMessagebitf_dist->net_brdgSendValidationMessageprovisionerProvisionerbitf_dist->provisionerProvisionableData::Bitfieldbitf_dist->runt_apiRequest::Validatiorsbitf_dist->runt_apiRequest::SessionIndexForChildbitf_signBitfield Signingbitf_sign->av_storeQueryChunkAvailabilitybitf_sign->bitf_distDistributeBitfieldbitf_sign->runt_apiRequest::AvailabilityCorescand_backCandidate Backingcand_back->av_storeStoreAvailableDatacand_selCandidate Selectioncand_back->cand_selInvalidcand_valCandidate Validationcand_back->cand_valValidateFromChainStatepov_distPoV Distributioncand_back->pov_distFetchPoVcand_back->pov_distDistributePoVcand_back->provisionerProvisionableData::MisbehaviorReportcand_back->provisionerProvisionableData::BackedCandidatestmt_distStatement Distributioncand_back->stmt_distSharecand_sel->cand_backSecondcoll_protCollator Protocolcand_sel->coll_protFetchCollationcand_sel->coll_protReportCollatorcand_val->runt_apiRequest::PersistedValidationDatacand_val->runt_apiRequest::ValidationCodecand_val->runt_apiRequest::CheckValidationOutputscoll_genCollation Generationcoll_gen->coll_protDistributeCollationcoll_prot->cand_selCollationcoll_prot->net_brdgReportPeercoll_prot->net_brdgDeclarecoll_prot->net_brdgAdvertiseCollationcoll_prot->net_brdgCollationcoll_prot->net_brdgRequestCollationnet_brdg->avail_distNetworkBridgeUpdatenet_brdg->bitf_distNetworkBridgeUpdatenet_brdg->coll_protNetworkBridgeUpdatenet_brdg->pov_distNetworkBridgeUpdatenet_brdg->stmt_distNetworkBridgeUpdatepov_dist->net_brdgSendValidationMessagepov_dist->net_brdgReportPeerprovisioner->cand_backGetBackedCandidatesprovisioner->chn_apiBlockNumberstmt_dist->cand_backStatementstmt_dist->net_brdgSendValidationMessagestmt_dist->net_brdgReportPeerstmt_dist->runt_apiRequest::Validatorsstmt_dist->runt_apiRequest::SessionIndexForChild

The Path to Inclusion (Node Side)

Let's contextualize that diagram a bit by following a parachain block from its creation through finalization. Parachains can use completely arbitrary processes to generate blocks. The relay chain doesn't know or care about the details; each parachain just needs to provide a collator.

Note: Inter-subsystem communications are relayed via the overseer, but that step is omitted here for brevity.

Note: Dashed lines indicate a request/response cycle, where the response is communicated asynchronously via a oneshot channel. Adjacent dashed lines may be processed in parallel.

sequenceDiagram
    participant Overseer
    participant CollationGeneration
    participant RuntimeApi
    participant CollatorProtocol

    Overseer ->> CollationGeneration: ActiveLeavesUpdate
    loop for each activated head
        CollationGeneration -->> RuntimeApi: Request availability cores
        CollationGeneration -->> RuntimeApi: Request validators

        Note over CollationGeneration: Determine an appropriate ScheduledCore <br/>and OccupiedCoreAssumption

        CollationGeneration -->> RuntimeApi: Request full validation data

        Note over CollationGeneration: Build the collation

        CollationGeneration ->> CollatorProtocol: DistributeCollation
    end

The DistributeCollation messages that CollationGeneration sends to the CollatorProtocol contains two items: a CandidateReceipt and PoV. The CollatorProtocol is then responsible for distributing that collation to interested validators. However, not all potential collations are of interest. The CandidateSelection subsystem is responsible for determining which collations are interesting, before CollatorProtocol actually fetches the collation.

sequenceDiagram
    participant CollationGeneration
    participant CS as CollatorProtocol::CollatorSide
    participant NB as NetworkBridge
    participant VS as CollatorProtocol::ValidatorSide
    participant CandidateSelection

    CollationGeneration ->> CS: DistributeCollation
    CS -->> NB: ConnectToValidators

    Note over CS,NB: This connects to multiple validators.

    CS ->> NB: Declare
    NB ->> VS: Declare

    Note over CS: Ensure that the connected validator is among<br/>the para's validator set. Otherwise, skip it.

    CS ->> NB: AdvertiseCollation
    NB ->> VS: AdvertiseCollation

    VS ->> CandidateSelection: Collation

    Note over CandidateSelection: Lots of other machinery in play here,<br/>but there are only three outcomes from the<br/>perspective of the `CollatorProtocol`:

    alt happy path
        CandidateSelection -->> VS: FetchCollation
        Activate VS
        VS ->> NB: RequestCollation
        NB ->> CS: RequestCollation
        CS ->> NB: Collation
        NB ->> VS: Collation
        Deactivate VS

    else collation invalid or unexpected
        CandidateSelection ->> VS: ReportCollator
        VS ->> NB: ReportPeer

    else CandidateSelection already selected a different candidate
        Note over CandidateSelection: silently drop
    end

Assuming we hit the happy path, flow continues with CandidateSelection receiving a (candidate_receipt, pov) as the return value from its FetchCollation request. The only time CandidateSelection actively requests a collation is when it hasn't yet seconded one for some relay_parent, and is ready to second.

sequenceDiagram
    participant CS as CandidateSelection
    participant CB as CandidateBacking
    participant CV as CandidateValidation
    participant PV as Provisioner
    participant SD as StatementDistribution
    participant PD as PoVDistribution

    CS ->> CB: Second
    % fn validate_and_make_available
    CB -->> CV: ValidateFromChainState

    Note over CB,CV: There's some complication in the source, as<br/>candidates are actually validated in a separate task.

    alt valid
        Note over CB: This is where we transform the CandidateReceipt into a CommittedCandidateReceipt
        % CandidateBackingJob::sign_import_and_distribute_statement
        % CandidateBackingJob::import_statement
        CB ->> PV: ProvisionableData::BackedCandidate
        % CandidateBackingJob::issue_new_misbehaviors
        opt if there is misbehavior to report
            CB ->> PV: ProvisionableData::MisbehaviorReport
        end
        % CandidateBackingJob::distribute_signed_statement
        CB ->> SD: Share
        % CandidateBackingJob::distribute_pov
        CB ->> PD: DistributePoV
    else invalid
        CB ->> CS: Invalid
    end

At this point, you'll see that control flows in two directions: to StatementDistribution to distribute the SignedStatement, and to PoVDistribution to distribute the PoV. However, that's largely a mirage: while the initial implementation distributes PoVs by gossip, that's inefficient, and will be replaced with a system which fetches PoVs only when actually necessary.

TODO: figure out more precisely the current status and plans; write them up

Therefore, we'll follow the SignedStatement. The StatementDistribution subsystem is largely concerned with implementing a gossip protocol:

sequenceDiagram
    participant SD as StatementDistribution
    participant NB as NetworkBridge

    alt On receipt of a<br/>SignedStatement from CandidateBacking
        % fn circulate_statement_and_dependents
        SD ->> NB: SendValidationMessage

        Note right of NB: Bridge sends validation message to all appropriate peers
    else On receipt of peer validation message
        NB ->> SD: NetworkBridgeUpdate

        % fn handle_incoming_message
        alt if we aren't already aware of the relay parent for this statement
            SD ->> NB: ReportPeer
        end

        % fn circulate_statement
        opt if we know of peers who haven't seen this message, gossip it
            SD ->> NB: SendValidationMessage
        end
    end

But who are these Listeners who've asked to be notified about incoming SignedStatements? Nobody, as yet.

Let's pick back up with the PoV Distribution subsystem.

sequenceDiagram
    participant CB as CandidateBacking
    participant PD as PoVDistribution
    participant Listener
    participant NB as NetworkBridge

    CB ->> PD: DistributePoV

    Note over PD,Listener: Various subsystems can register listeners for when PoVs arrive

    loop for each Listener
        PD ->> Listener: Arc<PoV>
    end

    Note over PD: Gossip to connected peers

    PD ->> NB: SendPoV

    Note over PD,NB: On receipt of a network PoV, PovDistribution forwards it to each Listener.<br/>It also penalizes bad gossipers.

Unlike in the case of StatementDistribution, there is another subsystem which in various circumstances already registers a listener to be notified when a new PoV arrives: CandidateBacking. Note that this is the second time that CandidateBacking has gotten involved. The first instance was from the perspective of the validator choosing to second a candidate via its CandidateSelection subsystem. This time, it's from the perspective of some other validator, being informed that this foreign PoV has been received.

sequenceDiagram
    participant SD as StatementDistribution
    participant CB as CandidateBacking
    participant PD as PoVDistribution
    participant AS as AvailabilityStore

    SD ->> CB: Statement
    % CB::maybe_validate_and_import => CB::kick_off_validation_work
    CB -->> PD: FetchPoV
    Note over CB,PD: This call creates the Listener from the previous diagram

    CB ->> AS: StoreAvailableData

At this point, things have gone a bit nonlinear. Let's pick up the thread again with BitfieldSigning. As the Overseer activates each relay parent, it starts a BitfieldSigningJob which operates on an extremely simple metric: after creation, it immediately goes to sleep for 1.5 seconds. On waking, it records the state of the world pertaining to availability at that moment.

sequenceDiagram
    participant OS as Overseer
    participant BS as BitfieldSigning
    participant RA as RuntimeApi
    participant AS as AvailabilityStore
    participant BD as BitfieldDistribution

    OS ->> BS: ActiveLeavesUpdate
    loop for each activated relay parent
        Note over BS: Wait 1.5 seconds
        BS -->> RA: Request::AvailabilityCores
        loop for each availability core
            BS -->> AS: QueryChunkAvailability
        end
        BS ->> BD: DistributeBitfield
    end

BitfieldDistribution is, like the other *Distribution subsystems, primarily interested in implementing a peer-to-peer gossip network propagating its particular messages. However, it also serves as an essential relay passing the message along.

sequenceDiagram
    participant BS as BitfieldSigning
    participant BD as BitfieldDistribution
    participant NB as NetworkBridge
    participant PV as Provisioner

    BS ->> BD: DistributeBitfield
    BD ->> PV: ProvisionableData::Bitfield
    BD ->> NB: SendValidationMessage::BitfieldDistribution::Bitfield

We've now seen the message flow to the Provisioner: both CandidateBacking and BitfieldDistribution contribute provisionable data. Now, let's look at that subsystem.

Much like the BitfieldSigning subsystem, the Provisioner creates a new job for each newly-activated leaf, and starts a timer. Unlike BitfieldSigning, we won't depict that part of the process, because the Provisioner also has other things going on.

sequenceDiagram
    participant A as Arbitrary
    participant PV as Provisioner
    participant CB as CandidateBacking
    participant BD as BitfieldDistribution
    participant RA as RuntimeApi
    participant PI as ParachainsInherentDataProvider

    alt receive provisionable data
        alt
            CB ->> PV: ProvisionableData
        else
            BD ->> PV: ProvisionableData
        end

        loop over stored Senders
            PV ->> A: ProvisionableData
        end

        Note over PV: store bitfields and backed candidates
    else receive request for inherent data
        PI ->> PV: RequestInherentData
        alt we have already constructed the inherent data
            PV ->> PI: send the inherent data
        else we have not yet constructed the inherent data
            Note over PV,PI: Store the return sender without sending immediately
        end
    else timer times out
        note over PV: Waited 2 seconds
        PV -->> RA: RuntimeApiRequest::AvailabilityCores
        Note over PV: construct and store the inherent data
        loop over stored inherent data requests
            PV ->> PI: (SignedAvailabilityBitfields, BackedCandidates)
        end
    end

In principle, any arbitrary subsystem could send a RequestInherentData to the Provisioner. In practice, only the ParachainsInherentDataProvider does so.

The tuple (SignedAvailabilityBitfields, BackedCandidates, ParentHeader) is injected by the ParachainsInherentDataProvider into the inherent data. From that point on, control passes from the node to the runtime.

Overseer

The overseer is responsible for these tasks:

  1. Setting up, monitoring, and handing failure for overseen subsystems.
  2. Providing a "heartbeat" of which relay-parents subsystems should be working on.
  3. Acting as a message bus between subsystems.

The hierarchy of subsystems:

+--------------+      +------------------+    +--------------------+
|              |      |                  |---->   Subsystem A      |
| Block Import |      |                  |    +--------------------+
|    Events    |------>                  |    +--------------------+
+--------------+      |                  |---->   Subsystem B      |
                      |   Overseer       |    +--------------------+
+--------------+      |                  |    +--------------------+
|              |      |                  |---->   Subsystem C      |
| Finalization |------>                  |    +--------------------+
|    Events    |      |                  |    +--------------------+
|              |      |                  |---->   Subsystem D      |
+--------------+      +------------------+    +--------------------+

The overseer determines work to do based on block import events and block finalization events. It does this by keeping track of the set of relay-parents for which work is currently being done. This is known as the "active leaves" set. It determines an initial set of active leaves on startup based on the data on-disk, and uses events about blockchain import to update the active leaves. Updates lead to OverseerSignal::ActiveLeavesUpdate being sent according to new relay-parents, as well as relay-parents to stop considering. Block import events inform the overseer of leaves that no longer need to be built on, now that they have children, and inform us to begin building on those children. Block finalization events inform us when we can stop focusing on blocks that appear to have been orphaned.

The overseer is also responsible for tracking the freshness of active leaves. Leaves are fresh when they're encountered for the first time, and stale when they're encountered for subsequent times. This can occur after chain reversions or when the fork-choice rule abandons some chain. This distinction is used to manage Reversion Safety. Consensus messages are often localized to a specific relay-parent, and it is often a misbehavior to equivocate or sign two conflicting messages. When reverting the chain, we may begin work on a leaf that subsystems have already signed messages for. Subsystems which need to account for reversion safety should avoid performing work on stale leaves.

The overseer's logic can be described with these functions:

On Startup

  • Start all subsystems
  • Determine all blocks of the blockchain that should be built on. This should typically be the head of the best fork of the chain we are aware of. Sometimes add recent forks as well.
  • Send an OverseerSignal::ActiveLeavesUpdate to all subsystems with activated containing each of these blocks.
  • Begin listening for block import and finality events

On Block Import Event

  • Apply the block import event to the active leaves. A new block should lead to its addition to the active leaves set and its parent being deactivated.
  • Mark any stale leaves as stale. The overseer should track all leaves it activates to determine whether leaves are fresh or stale.
  • Send an OverseerSignal::ActiveLeavesUpdate message to all subsystems containing all activated and deactivated leaves.
  • Ensure all ActiveLeavesUpdate messages are flushed before resuming activity as a message router.

TODO: in the future, we may want to avoid building on too many sibling blocks at once. the notion of a "preferred head" among many competing sibling blocks would imply changes in our "active leaves" update rules here

On Finalization Event

  • Note the height h of the newly finalized block B.
  • Prune all leaves from the active leaves which have height <= h and are not B.
  • Issue OverseerSignal::ActiveLeavesUpdate containing all deactivated leaves.

On Subsystem Failure

Subsystems are essential tasks meant to run as long as the node does. Subsystems can spawn ephemeral work in the form of jobs, but the subsystems themselves should not go down. If a subsystem goes down, it will be because of a critical error that should take the entire node down as well.

Communication Between Subsystems

When a subsystem wants to communicate with another subsystem, or, more typically, a job within a subsystem wants to communicate with its counterpart under another subsystem, that communication must happen via the overseer. Consider this example where a job on subsystem A wants to send a message to its counterpart under subsystem B. This is a realistic scenario, where you can imagine that both jobs correspond to work under the same relay-parent.

     +--------+                                                           +--------+
     |        |                                                           |        |
     |Job A-1 | (sends message)                       (receives message)  |Job B-1 |
     |        |                                                           |        |
     +----|---+                                                           +----^---+
          |                  +------------------------------+                  ^
          v                  |                              |                  |
+---------v---------+        |                              |        +---------|---------+
|                   |        |                              |        |                   |
| Subsystem A       |        |       Overseer / Message     |        | Subsystem B       |
|                   -------->>                  Bus         -------->>                   |
|                   |        |                              |        |                   |
+-------------------+        |                              |        +-------------------+
                             |                              |
                             +------------------------------+

First, the subsystem that spawned a job is responsible for handling the first step of the communication. The overseer is not aware of the hierarchy of tasks within any given subsystem and is only responsible for subsystem-to-subsystem communication. So the sending subsystem must pass on the message via the overseer to the receiving subsystem, in such a way that the receiving subsystem can further address the communication to one of its internal tasks, if necessary.

This communication prevents a certain class of race conditions. When the Overseer determines that it is time for subsystems to begin working on top of a particular relay-parent, it will dispatch a ActiveLeavesUpdate message to all subsystems to do so, and those messages will be handled asynchronously by those subsystems. Some subsystems will receive those messsages before others, and it is important that a message sent by subsystem A after receiving ActiveLeavesUpdate message will arrive at subsystem B after its ActiveLeavesUpdate message. If subsystem A maintained an independent channel with subsystem B to communicate, it would be possible for subsystem B to handle the side message before the ActiveLeavesUpdate message, but it wouldn't have any logical course of action to take with the side message - leading to it being discarded or improperly handled. Well-architectured state machines should have a single source of inputs, so that is what we do here.

One exception is reasonable to make for responses to requests. A request should be made via the overseer in order to ensure that it arrives after any relevant ActiveLeavesUpdate message. A subsystem issuing a request as a result of a ActiveLeavesUpdate message can safely receive the response via a side-channel for two reasons:

  1. It's impossible for a request to be answered before it arrives, it is provable that any response to a request obeys the same ordering constraint.
  2. The request was sent as a result of handling a ActiveLeavesUpdate message. Then there is no possible future in which the ActiveLeavesUpdate message has not been handled upon the receipt of the response.

So as a single exception to the rule that all communication must happen via the overseer we allow the receipt of responses to requests via a side-channel, which may be established for that purpose. This simplifies any cases where the outside world desires to make a request to a subsystem, as the outside world can then establish a side-channel to receive the response on.

It's important to note that the overseer is not aware of the internals of subsystems, and this extends to the jobs that they spawn. The overseer isn't aware of the existence or definition of those jobs, and is only aware of the outer subsystems with which it interacts. This gives subsystem implementations leeway to define internal jobs as they see fit, and to wrap a more complex hierarchy of state machines than having a single layer of jobs for relay-parent-based work. Likewise, subsystems aren't required to spawn jobs. Certain types of subsystems, such as those for shared storage or networking resources, won't perform block-based work but would still benefit from being on the Overseer's message bus. These subsystems can just ignore the overseer's signals for block-based work.

Furthermore, the protocols by which subsystems communicate with each other should be well-defined irrespective of the implementation of the subsystem. In other words, their interface should be distinct from their implementation. This will prevent subsystems from accessing aspects of each other that are beyond the scope of the communication boundary.

On shutdown

Send an OverseerSignal::Conclude message to each subsystem and wait some time for them to conclude before hard-exiting.

GRANDPA Voting Rule

Specifics on the motivation and types of constraints we apply to the GRANDPA voting logic as well as the definitions of viable and finalizable blocks can be found in the Chain Selection Protocol section. The subsystem which provides us with viable leaves is the Chain Selection Subsystem.

GRANDPA's regular voting rule is for each validator to select the longest chain they are aware of. GRANDPA proceeds in rounds, collecting information from all online validators and determines the blocks that a supermajority of validators all have in common with each other.

The low-level GRANDPA logic will provide us with a required block. We can find the best leaf containing that block in its chain with the ChainSelectionMessage::BestLeafContaining. If the result is None, then we will simply cast a vote on the required block.

The viable leaves provided from the chain selection subsystem are not necessarily finalizable, so we need to perform further work to discover the finalizable ancestor of the block. The first constraint is to avoid voting on any unapproved block. The highest approved ancestor of a given block can be determined by querying the Approval Voting subsystem via the ApprovalVotingMessage::ApprovedAncestor message. If the response is Some, we continue and apply the second constraint. The second constraint is to avoid voting on any block containing a candidate undergoing an active dispute. The list of block hashes and candidates returned from ApprovedAncestor should be reversed, and passed to the DisputeCoordinatorMessage::DetermineUndisputedChain to determine the finalizable block which will be our eventual vote.

Collators

Collators are special nodes which bridge a parachain to the relay chain. They are simultaneously full nodes of the parachain, and at least light clients of the relay chain. Their overall contribution to the system is the generation of Proofs of Validity for parachain candidates.

The Collation Generation subsystem triggers collators to produce collations and then forwards them to Collator Protocol to circulate to validators.

Collation Generation

The collation generation subsystem is executed on collator nodes and produces candidates to be distributed to validators. If configured to produce collations for a para, it produces collations and then feeds them to the Collator Protocol subsystem, which handles the networking.

Protocol

Collation generation for Parachains currently works in the following way:

  1. A new relay chain block is imported.
  2. The collation generation subsystem checks if the core associated to the parachain is free and if yes, continues.
  3. Collation generation calls our collator callback, if present, to generate a PoV. If none exists, do nothing.
  4. Authoring logic determines if the current node should build a PoV.
  5. Build new PoV and give it back to collation generation.

Messages

Incoming

  • ActiveLeaves
    • Notification of a change in the set of active leaves.
    • Triggers collation generation procedure outlined in "Protocol" section.
  • CollationGenerationMessage::Initialize
    • Initializes the subsystem. Carries a config.
    • No more than one initialization message should ever be sent to the collation generation subsystem.
    • Sent by a collator to initialize this subsystem.
  • CollationGenerationMessage::SubmitCollation
    • If the subsystem isn't initialized or the relay-parent is too old to be relevant, ignore the message.
    • Otherwise, use the provided parameters to generate a [CommittedCandidateReceipt]
    • Submit the collation to the collator-protocol with CollatorProtocolMessage::DistributeCollation.

Outgoing

  • CollatorProtocolMessage::DistributeCollation
    • Provides a generated collation to distribute to validators.

Functionality

The process of generating a collation for a parachain is very parachain-specific. As such, the details of how to do so are left beyond the scope of this description. The subsystem should be implemented as an abstract wrapper, which is aware of this configuration:

#![allow(unused)]
fn main() {
/// The output of a collator.
///
/// This differs from `CandidateCommitments` in two ways:
///
/// - does not contain the erasure root; that's computed at the Polkadot level, not at Cumulus
/// - contains a proof of validity.
pub struct Collation {
  /// Messages destined to be interpreted by the Relay chain itself.
  pub upward_messages: Vec<UpwardMessage>,
  /// The horizontal messages sent by the parachain.
  pub horizontal_messages: Vec<OutboundHrmpMessage<ParaId>>,
  /// New validation code.
  pub new_validation_code: Option<ValidationCode>,
  /// The head-data produced as a result of execution.
  pub head_data: HeadData,
  /// Proof to verify the state transition of the parachain.
  pub proof_of_validity: PoV,
  /// The number of messages processed from the DMQ.
  pub processed_downward_messages: u32,
  /// The mark which specifies the block number up to which all inbound HRMP messages are processed.
  pub hrmp_watermark: BlockNumber,
}

/// Result of the [`CollatorFn`] invocation.
pub struct CollationResult {
  /// The collation that was build.
  pub collation: Collation,
  /// An optional result sender that should be informed about a successfully seconded collation.
  ///
  /// There is no guarantee that this sender is informed ever about any result, it is completely okay to just drop it.
  /// However, if it is called, it should be called with the signed statement of a parachain validator seconding the
  /// collation.
  pub result_sender: Option<oneshot::Sender<CollationSecondedSignal>>,
}

/// Signal that is being returned when a collation was seconded by a validator.
pub struct CollationSecondedSignal {
  /// The hash of the relay chain block that was used as context to sign [`Self::statement`].
  pub relay_parent: Hash,
  /// The statement about seconding the collation.
  ///
  /// Anything else than `Statement::Seconded` is forbidden here.
  pub statement: SignedFullStatement,
}

/// Collation function.
///
/// Will be called with the hash of the relay chain block the parachain block should be build on and the
/// [`ValidationData`] that provides information about the state of the parachain on the relay chain.
///
/// Returns an optional [`CollationResult`].
pub type CollatorFn = Box<
  dyn Fn(
      Hash,
      &PersistedValidationData,
    ) -> Pin<Box<dyn Future<Output = Option<CollationResult>> + Send>>
    + Send
    + Sync,
>;

/// Configuration for the collation generator
pub struct CollationGenerationConfig {
  /// Collator's authentication key, so it can sign things.
  pub key: CollatorPair,
  /// Collation function. See [`CollatorFn`] for more details.
  pub collator: Option<CollatorFn>,
  /// The parachain that this collator collates for
  pub para_id: ParaId,
}
}

The configuration should be optional, to allow for the case where the node is not run with the capability to collate.

Summary in plain English

  • Collation (output of a collator)

    • Contains the PoV (proof to verify the state transition of the parachain) and other data.
  • Collation result

    • Contains the collation, and an optional result sender for a collation-seconded signal.
  • Collation seconded signal

    • The signal that is returned when a collation was seconded by a validator.
  • Collation function

    • Called with the relay chain block the parablock will be built on top of.
    • Called with the validation data.
      • Provides information about the state of the parachain on the relay chain.
  • Collation generation config

    • Contains collator's authentication key, optional collator function, and parachain ID.

Collator Protocol

The Collator Protocol implements the network protocol by which collators and validators communicate. It is used by collators to distribute collations to validators and used by validators to accept collations by collators.

Collator-to-Validator networking is more difficult than Validator-to-Validator networking because the set of possible collators for any given para is unbounded, unlike the validator set. Validator-to-Validator networking protocols can easily be implemented as gossip because the data can be bounded, and validators can authenticate each other by their PeerIds for the purposes of instantiating and accepting connections.

Since, at least at the level of the para abstraction, the collator-set for any given para is unbounded, validators need to make sure that they are receiving connections from capable and honest collators and that their bandwidth and time are not being wasted by attackers. Communicating across this trust-boundary is the most difficult part of this subsystem.

Validation of candidates is a heavy task, and furthermore, the PoV itself is a large piece of data. Empirically, PoVs are on the order of 10MB.

TODO: note the incremental validation function Ximin proposes at https://github.com/paritytech/polkadot/issues/1348

As this network protocol serves as a bridge between collators and validators, it communicates primarily with one subsystem on behalf of each. As a collator, this will receive messages from the CollationGeneration subsystem. As a validator, this will communicate only with the CandidateBacking.

Protocol

Input: CollatorProtocolMessage

Output:

Functionality

This network protocol uses the Collation peer-set of the NetworkBridge.

It uses the CollatorProtocolV1Message as its WireMessage

Since this protocol functions both for validators and collators, it is easiest to go through the protocol actions for each of them separately.

Validators and collators.

%3c1Collator 1v1Validator 1c1->v1v2Validator 2c1->v2c2Collator 2c2->v2

Collators

It is assumed that collators are only collating on a single parachain. Collations are generated by the Collation Generation subsystem. We will keep up to one local collation per relay-parent, based on DistributeCollation messages. If the para is not scheduled on any core, at the relay-parent, or the relay-parent isn't in the active-leaves set, we ignore the message as it must be invalid in that case - although this indicates a logic error elsewhere in the node.

We keep track of the Para ID we are collating on as a collator. This starts as None, and is updated with each CollateOn message received. If the ParaId of a collation requested to be distributed does not match the one we expect, we ignore the message.

As with most other subsystems, we track the active leaves set by following ActiveLeavesUpdate signals.

For the purposes of actually distributing a collation, we need to be connected to the validators who are interested in collations on that ParaId at this point in time. We assume that there is a discovery API for connecting to a set of validators.

As seen in the Scheduler Module of the runtime, validator groups are fixed for an entire session and their rotations across cores are predictable. Collators will want to do these things when attempting to distribute collations at a given relay-parent:

  • Determine which core the para collated-on is assigned to.
  • Determine the group on that core.
  • Issue a discovery request for the validators of the current group withNetworkBridgeMessage::ConnectToValidators.

Once connected to the relevant peers for the current group assigned to the core (transitively, the para), advertise the collation to any of them which advertise the relay-parent in their view (as provided by the Network Bridge). If any respond with a request for the full collation, provide it. However, we only send one collation at a time per relay parent, other requests need to wait. This is done to reduce the bandwidth requirements of a collator and also increases the chance to fully send the collation to at least one validator. From the point where one validator has received the collation and seconded it, it will also start to share this collation with other validators in its backing group. Upon receiving a view update from any of these peers which includes a relay-parent for which we have a collation that they will find relevant, advertise the collation to them if we haven't already.

Validators

On the validator side of the protocol, validators need to accept incoming connections from collators. They should keep some peer slots open for accepting new speculative connections from collators and should disconnect from collators who are not relevant.

GDeclaring, advertising, and providing collationscluster_collatorCollatorcluster_validatorValidatorc1v1c1->v1Declare and advertisec2v2c2->v2Providev1->c2Requestv2->v2Note Good/Bad

When peers connect to us, they can Declare that they represent a collator with given public key and intend to collate on a specific para ID. Once they've declared that, and we checked their signature, they can begin to send advertisements of collations. The peers should not send us any advertisements for collations that are on a relay-parent outside of our view or for a para outside of the one they've declared.

The protocol tracks advertisements received and the source of the advertisement. The advertisement source is the PeerId of the peer who sent the message. We accept one advertisement per collator per source per relay-parent.

As a validator, we will handle requests from other subsystems to fetch a collation on a specific ParaId and relay-parent. These requests are made with the request response protocol CollationFetchingRequest request. To do so, we need to first check if we have already gathered a collation on that ParaId and relay-parent. If not, we need to select one of the advertisements and issue a request for it. If we've already issued a request, we shouldn't issue another one until the first has returned.

When acting on an advertisement, we issue a Requests::CollationFetchingV1. However, we only request one collation at a time per relay parent. This reduces the bandwidth requirements and as we can second only one candidate per relay parent, the others are probably not required anyway. If the request times out, we need to note the collator as being unreliable and reduce its priority relative to other collators.

As a validator, once the collation has been fetched some other subsystem will inspect and do deeper validation of the collation. The subsystem will report to this subsystem with a CollatorProtocolMessage::ReportCollator. In that case, if we are connected directly to the collator, we apply a cost to the PeerId associated with the collator and potentially disconnect or blacklist it. If the collation is seconded, we notify the collator and apply a benefit to the PeerId associated with the collator.

Interaction with Candidate Backing

As collators advertise the availability, a validator will simply second the first valid parablock candidate per relay head by sending a CandidateBackingMessage::Second. Note that this message contains the relay parent of the advertised collation, the candidate receipt and the PoV.

Subsequently, once a valid parablock candidate has been seconded, the CandidateBacking subsystem will send a CollatorProtocolMessage::Seconded, which will trigger this subsystem to notify the collator at the PeerId that first advertised the parablock on the seconded relay head of their successful seconding.

Future Work

Several approaches have been discussed, but all have some issues:

  • The current approach is very straightforward. However, that protocol is vulnerable to a single collator which, as an attack or simply through chance, gets its block candidate to the node more often than its fair share of the time.
  • If collators produce blocks via Aura, BABE or in future Sassafras, it may be possible to choose an "Official" collator for the round, but it may be tricky to ensure that the PVF logic is enforced at collator leader election.
  • We could use relay-chain BABE randomness to generate some delay D on the order of 1 second, +- 1 second. The collator would then second the first valid parablock which arrives after D, or in case none has arrived by 2*D, the last valid parablock which has arrived. This makes it very hard for a collator to game the system to always get its block nominated, but it reduces the maximum throughput of the system by introducing delay into an already tight schedule.
  • A variation of that scheme would be to have a fixed acceptance window D for parablock candidates and keep track of count C: the number of parablock candidates received. At the end of the period D, we choose a random number I in the range [0, C) and second the block at Index I. Its drawback is the same: it must wait the full D period before seconding any of its received candidates, reducing throughput.
  • In order to protect against DoS attacks, it may be prudent to run throw out collations from collators that have behaved poorly (whether recently or historically) and subsequently only verify the PoV for the most suitable of collations.

Backing Subsystems

The backing subsystems, when conceived as a black box, receive an arbitrary quantity of parablock candidates and associated proofs of validity from arbitrary untrusted collators. From these, they produce a bounded quantity of backable candidates which relay chain block authors may choose to include in a subsequent block.

In broad strokes, the flow operates like this:

  • Candidate Selection winnows the field of parablock candidates, selecting up to one of them to second.
  • Candidate Backing ensures that a seconding candidate is valid, then generates the appropriate Statement. It also keeps track of which candidates have received the backing of a quorum of other validators.
  • Statement Distribution is the networking component which ensures that all validators receive each others' statements.
  • PoV Distribution is the networking component which ensures that validators considering a candidate can get the appropriate PoV.

Candidate Backing

The Candidate Backing subsystem ensures every parablock considered for relay block inclusion has been seconded by at least one validator, and approved by a quorum. Parablocks for which not enough validators will assert correctness are discarded. If the block later proves invalid, the initial backers are slashable; this gives polkadot a rational threat model during subsequent stages.

Its role is to produce backable candidates for inclusion in new relay-chain blocks. It does so by issuing signed Statements and tracking received statements signed by other validators. Once enough statements are received, they can be combined into backing for specific candidates.

Note that though the candidate backing subsystem attempts to produce as many backable candidates as possible, it does not attempt to choose a single authoritative one. The choice of which actually gets included is ultimately up to the block author, by whatever metrics it may use; those are opaque to this subsystem.

Once a sufficient quorum has agreed that a candidate is valid, this subsystem notifies the Provisioner, which in turn engages block production mechanisms to include the parablock.

Protocol

Input: CandidateBackingMessage

Output:

Functionality

The Collator Protocol subsystem is the primary source of non-overseer messages into this subsystem. That subsystem generates appropriate CandidateBackingMessages and passes them to this subsystem.

This subsystem requests validation from the Candidate Validation and generates an appropriate Statement. All Statements are then passed on to the Statement Distribution subsystem to be gossiped to peers. When Candidate Validation decides that a candidate is invalid, and it was recommended to us to second by our own Collator Protocol subsystem, a message is sent to the Collator Protocol subsystem with the candidate's hash so that the collator which recommended it can be penalized.

The subsystem should maintain a set of handles to Candidate Backing Jobs that are currently live, as well as the relay-parent to which they correspond.

On Overseer Signal

  • If the signal is an OverseerSignal::ActiveLeavesUpdate:
    • spawn a Candidate Backing Job for each activated head referring to a fresh leaf, storing a bidirectional channel with the Candidate Backing Job in the set of handles.
    • cease the Candidate Backing Job for each deactivated head, if any.
  • If the signal is an OverseerSignal::Conclude: Forward conclude messages to all jobs, wait a small amount of time for them to join, and then exit.

On Receiving CandidateBackingMessage

  • If the message is a CandidateBackingMessage::GetBackedCandidates, get all backable candidates from the statement table and send them back.
  • If the message is a CandidateBackingMessage::Second, sign and dispatch a Seconded statement only if we have not seconded any other candidate and have not signed a Valid statement for the requested candidate. Signing both a Seconded and Valid message is a double-voting misbehavior with a heavy penalty, and this could occur if another validator has seconded the same candidate and we've received their message before the internal seconding request.
  • If the message is a CandidateBackingMessage::Statement, count the statement to the quorum. If the statement in the message is Seconded and it contains a candidate that belongs to our assignment, request the corresponding PoV from the backing node via AvailabilityDistribution and launch validation. Issue our own Valid or Invalid statement as a result.

If the seconding node did not provide us with the PoV we will retry fetching from other backing validators.

big TODO: "contextual execution"

  • At the moment we only allow inclusion of new parachain candidates validated by current validators.
  • Allow inclusion of old parachain candidates validated by current validators.
  • Allow inclusion of old parachain candidates validated by old validators.

This will probably blur the lines between jobs, will probably require inter-job communication and a short-term memory of recently backable, but not backed candidates.

Candidate Backing Job

The Candidate Backing Job represents the work a node does for backing candidates with respect to a particular relay-parent.

The goal of a Candidate Backing Job is to produce as many backable candidates as possible. This is done via signed Statements by validators. If a candidate receives a majority of supporting Statements from the Parachain Validators currently assigned, then that candidate is considered backable.

On Startup

On Receiving New Candidate Backing Message

#![allow(unused)]
fn main() {
match msg {
  GetBackedCandidates(hashes, tx) => {
    // Send back a set of backable candidates.
  }
  CandidateBackingMessage::Second(hash, candidate) => {
    if candidate is unknown and in local assignment {
      if spawn_validation_work(candidate, parachain head, validation function).await == Valid {
        send(DistributePoV(pov))
      }
    }
  }
  CandidateBackingMessage::Statement(hash, statement) => {
    // count to the votes on this candidate
    if let Statement::Seconded(candidate) = statement {
      if candidate.parachain_id == our_assignment {
        spawn_validation_work(candidate, parachain head, validation function)
      }
    }
  }
}
}

Add Seconded statements and Valid statements to a quorum. If the quorum reaches a pre-defined threshold, send a ProvisionerMessage::ProvisionableData(ProvisionableData::BackedCandidate(CandidateReceipt)) message. Invalid statements that conflict with already witnessed Seconded and Valid statements for the given candidate, statements that are double-votes, self-contradictions and so on, should result in issuing a ProvisionerMessage::MisbehaviorReport message for each newly detected case of this kind.

Backing does not need to concern itself with providing statements to the dispute coordinator as the dispute coordinator scrapes them from chain. This way the import is batched and contains only statements that actually made it on some chain.

Validating Candidates.

#![allow(unused)]
fn main() {
fn spawn_validation_work(candidate, parachain head, validation function) {
  asynchronously {
    let pov = (fetch pov block).await

    let valid = (validate pov block).await;
    if valid {
      // make PoV available for later distribution. Send data to the availability store to keep.
      // sign and dispatch `valid` statement to network if we have not seconded the given candidate.
    } else {
      // sign and dispatch `invalid` statement to network.
    }
  }
}
}

Fetch PoV Block

Create a (sender, receiver) pair. Dispatch a AvailabilityDistributionMessage::FetchPoV{ validator_index, pov_hash, candidate_hash, tx, } and listen on the passed receiver for a response. Availability distribution will send the request to the validator specified by validator_index`, which might not be serving it for whatever reasons, therefore we need to retry with other backing validators in that case.

Validate PoV Block

Create a (sender, receiver) pair. Dispatch a CandidateValidationMessage::Validate(validation function, candidate, pov, BACKING_EXECUTION_TIMEOUT, sender) and listen on the receiver for a response.

Distribute Signed Statement

Dispatch a StatementDistributionMessage::Share(relay_parent, SignedFullStatementWithPVD).

Prospective Parachains

Overview

Purpose: Tracks and handles prospective parachain fragments and informs other backing-stage subsystems of work to be done.

"prospective":

  • [prə'spɛktɪv] adj.
  • future, likely, potential

Asynchronous backing changes the runtime to accept parachain candidates from a certain allowed range of historic relay-parents. This means we can now build prospective parachains – that is, trees of potential (but likely) future parachain blocks. This is the subsystem responsible for doing so.

Other subsystems such as Backing rely on Prospective Parachains, e.g. for determining if a candidate can be seconded. This subsystem is the main coordinator of work within the node for the collation and backing phases of parachain consensus.

Prospective Parachains is primarily an implementation of fragment trees. It also handles concerns such as:

  • the relay-chain being forkful
  • session changes

See the following sections for more details.

Fragment Trees

This subsystem builds up fragment trees, which are trees of prospective para candidates. Each path through the tree represents a possible state transition path for the para. Each potential candidate is a fragment, or a node, in the tree. Candidates are validated against constraints as they are added.

This subsystem builds up trees for each relay-chain block in the view, for each para. These fragment trees are used for:

  • providing backable candidates to other subsystems
  • sanity-checking that candidates can be seconded
  • getting seconded candidates under active leaves
  • etc.

For example, here is a tree with several possible paths:

Para Head registered by the relay chain:     included_head
                                                  ↲  ↳
depth 0:                                  head_0_a    head_0_b
                                             ↲            ↳
depth 1:                             head_1_a              head_1_b
                                  ↲      |     ↳
depth 2:                 head_2_a1   head_2_a2  head_2_a3

The Relay-Chain Being Forkful

We account for the same candidate possibly appearing in different forks. While we still build fragment trees for each head in each fork, we are efficient with how we reference candidates to save space.

Session Changes

Allowed ancestry doesn't cross session boundary. That is, you can only build on top of the freshest relay parent when the session starts. This is a current limitation that may be lifted in the future.

Also, runtime configuration values needed for constraints (such as max_pov_size) are constant within a session. This is important when building prospective validation data. This is unlikely to change.

Messages

Incoming

  • ActiveLeaves
    • Notification of a change in the set of active leaves.
    • Constructs fragment trees for each para for each new leaf.
  • ProspectiveParachainsMessage::IntroduceCandidate
    • Informs the subsystem of a new candidate.
    • Sent by the Backing Subsystem when it is importing a statement for a new candidate.
  • ProspectiveParachainsMessage::CandidateSeconded
    • Informs the subsystem that a previously introduced candidate has been seconded.
    • Sent by the Backing Subsystem when it is importing a statement for a new candidate after it sends IntroduceCandidate, if that wasn't rejected by Prospective Parachains.
  • ProspectiveParachainsMessage::CandidateBacked
    • Informs the subsystem that a previously introduced candidate has been backed.
    • Sent by the Backing Subsystem after it successfully imports a statement giving a candidate the necessary quorum of backing votes.
  • ProspectiveParachainsMessage::GetBackableCandidate
    • Get a backable candidate hash along with its relay parent for a given parachain, under a given relay-parent (leaf) hash, which is a descendant of given candidate hashes.
    • Sent by the Provisioner when requesting backable candidates, when selecting candidates for a given relay-parent.
  • ProspectiveParachainsMessage::GetHypotheticalFrontier
    • Gets the hypothetical frontier membership of candidates with the given properties under the specified active leaves' fragment trees.
    • Sent by the Backing Subsystem when sanity-checking whether a candidate can be seconded based on its hypothetical frontiers.
  • ProspectiveParachainsMessage::GetTreeMembership
    • Gets the membership of the candidate in all fragment trees.
    • Sent by the Backing Subsystem when it needs to update the candidates seconded at various depths under new active leaves.
  • ProspectiveParachainsMessage::GetMinimumRelayParents
    • Gets the minimum accepted relay-parent number for each para in the fragment tree for the given relay-chain block hash.
    • That is, this returns the minimum relay-parent block number in the same branch of the relay-chain which is accepted in the fragment tree for each para-id.
    • Sent by the Backing, Statement Distribution, and Collator Protocol subsystems when activating leaves in the implicit view.
  • ProspectiveParachainsMessage::GetProspectiveValidationData
    • Gets the validation data of some prospective candidate. The candidate doesn't need to be part of any fragment tree.
    • Sent by the Collator Protocol subsystem (validator side) when handling a fetched collation result.

Outgoing

  • RuntimeApiRequest::StagingParaBackingState
    • Gets the backing state of the given para (the constraints of the para and candidates pending availability).
  • RuntimeApiRequest::AvailabilityCores
    • Gets information on all availability cores.
  • ChainApiMessage::Ancestors
    • Requests the k ancestor block hashes of a block with the given hash.
  • ChainApiMessage::BlockHeader
    • Requests the block header by hash.

Glossary

  • Candidate storage: Stores candidates and information about them such as their relay-parents and their backing states. Is indexed in various ways.
  • Constraints:
    • Constraints on the actions that can be taken by a new parachain block.
    • Exhaustively define the set of valid inputs and outputs to parachain execution.
  • Fragment: A prospective para block (that is, a block not yet referenced by the relay-chain). Fragments are anchored to the relay-chain at a particular relay-parent.
  • Fragment tree:
    • A tree of fragments. Together, these fragments define one or more prospective paths a parachain's state may transition through.
    • See the "Fragment Tree" section.
  • Inclusion emulation: Emulation of the logic that the runtime uses for checking parachain blocks.
  • Relay-parent: A particular relay-chain block that a fragment is anchored to.
  • Scope: The scope of a fragment tree, defining limits on nodes within the tree.

Statement Distribution

This subsystem is responsible for distributing signed statements that we have generated and forwarding statements generated by our peers. Received candidate receipts and statements are passed to the Candidate Backing subsystem to handle producing local statements. On receiving StatementDistributionMessage::Share, this subsystem distributes the message across the network with redundency to ensure a fast backing process.

Overview

Goal: every well-connected node is aware of every next potential parachain block.

Validators can either:

  • receive parachain block from collator, check block, and gossip statement.
  • receive statements from other validators, check the parachain block if it originated within their own group, gossip forward statement if valid.

Validators must have statements, candidates, and persisted validation from all other validators. This is because we need to store statements from validators who've checked the candidate on the relay chain, so we know who to hold accountable in case of disputes. Any validator can be selected as the next relay-chain block author, and this is not revealed in advance for security reasons. As a result, all validators must have a up to date view of all possible parachain candidates + backing statements that could be placed on-chain in the next block.

This blog post puts it another way: "Validators who aren't assigned to the parachain still listen for the attestations [statements] because whichever validator ends up being the author of the relay-chain block needs to bundle up attested parachain blocks for several parachains and place them into the relay-chain block."

Backing-group quorum (that is, enough backing group votes) must be reached before the block author will consider the candidate. Therefore, validators need to consider all seconded candidates within their own group, because that's what they're assigned to work on. Validators only need to consider backable candidates from other groups. This informs the design of the statement distribution protocol to have separate phases for in-group and out-group distribution, respectively called "cluster" and "grid" mode (see below).

With Async Backing

Asynchronous backing changes the runtime to accept parachain candidates from a certain allowed range of historic relay-parents. These candidates must be backed by the group assigned to the parachain as-of their corresponding relay parents.

Protocol

To address the concern of dealing with large numbers of spam candidates or statements, the overall design approach is to combine a focused "clustering" protocol for legitimate fresh candidates with a broad-distribution "grid" protocol to quickly get backed candidates into the hands of many validators. Validators do not eagerly send each other heavy CommittedCandidateReceipt, but instead request these lazily through request/response protocols.

A high-level description of the protocol follows:

Messages

Nodes can send each other a few kinds of messages: Statement, BackedCandidateManifest, BackedCandidateAcknowledgement.

  • Statement messages contain only a signed compact statement, without full candidate info.
  • BackedCandidateManifest messages advertise a description of a backed candidate and stored statements.
  • BackedCandidateAcknowledgement messages acknowledge that a backed candidate is fully known.

Request/response protocol

Nodes can request the full CommittedCandidateReceipt and PersistedValidationData, along with statements, over a request/response protocol. This is the AttestedCandidateRequest; the response is AttestedCandidateResponse.

Importability and the Hypothetical Frontier

The prospective parachains subsystem maintains prospective "fragment trees" which can be used to determine whether a particular parachain candidate could possibly be included in the future. Candidates which either are within a fragment tree or would be part of a fragment tree if accepted are said to be in the "hypothetical frontier".

The statement-distribution subsystem keeps track of all candidates, and updates its knowledge of the hypothetical frontier based on events such as new relay parents, new confirmed candidates, and newly backed candidates.

We only consider statements as "importable" when the corresponding candidate is part of the hypothetical frontier, and only send "importable" statements to the backing subsystem itself.

Cluster Mode

  • Validator nodes are partitioned into groups (with some exceptions), and validators within a group at a relay-parent can send each other Statement messages for any candidates within that group and based on that relay-parent.
  • This is referred to as the "cluster" mode.
    • Right now these are the same as backing groups, though "cluster" specifically refers to the set of nodes communicating with each other in the first phase of distribution.
  • Seconded statements must be sent before Valid statements.
  • Seconded statements may only be sent to other members of the group when the candidate is fully known by the local validator.
    • "Fully known" means the validator has the full CommittedCandidateReceipt and PersistedValidationData, which it receives on request from other validators or from a collator.
    • The reason for this is that sending a statement (which is always a CompactStatement carrying nothing but a hash and signature) to the cluster, is also a signal that the sending node is available to request the candidate from.
    • This makes the protocol easier to reason about, while also reducing network messages about candidates that don't really exist.
  • Validators in a cluster receiving messages about unknown candidates request the candidate (and statements) from other cluster members which have it.
  • Spam considerations
    • The maximum depth of candidates allowed in asynchronous backing determines the maximum amount of Seconded statements originating from a validator V which each validator in a cluster may send to others. This bounds the number of candidates.
    • There is a small number of validators in each group, which further limits the amount of candidates.
  • We accept candidates which don't fit in the fragment trees of any relay parents.
    • "Accept" means "attempt to request and store in memory until useful or expired".
    • We listen to prospective parachains subsystem to learn of new additions to the fragment trees.
    • Use this to attempt to import the candidate later.

Grid Mode

  • Every consensus session provides randomness and a fixed validator set, which is used to build a redundant grid topology.
    • It's redundant in the sense that there are 2 paths from every node to every other node. See "Grid Topology" section for more details.
  • This grid topology is used to create a sending path from each validator group to every validator.
  • When a node observes a candidate as backed, it sends a BackedCandidateManifest to their "receiving" nodes.
  • If receiving nodes don't yet know the candidate, they request it.
  • Once they know the candidate, they respond with a BackedCandidateAcknowledgement.
  • Once two nodes perform a manifest/acknowledgement exchange, they can send Statement messages directly to each other for any new statements they might need.
    • This limits the amount of statements we'd have to deal with w.r.t. candidates that don't really exist. See "Manifest Exchange" section.
  • There are limitations on the number of candidates that can be advertised by each peer, similar to those in the cluster. Validators do not request candidates which exceed these limitations.
  • Validators request candidates as soon as they are advertised, but do not import the statements until the candidate is part of the hypothetical frontier, and do not re-advertise or acknowledge until the candidate is considered both backable and part of the hypothetical frontier.
  • Note that requesting is not an implicit acknowledgement, and an explicit acknowledgement must be sent upon receipt.

Messages

Incoming

  • ActiveLeaves
    • Notification of a change in the set of active leaves.
  • StatementDistributionMessage::Share
    • Notification of a locally-originating statement. That is, this statement comes from our node and should be distributed to other nodes.
    • Sent by the Backing Subsystem after it successfully imports a locally-originating statement.
  • StatementDistributionMessage::Backed
    • Notification of a candidate being backed (received enough validity votes from the backing group).
    • Sent by the Backing Subsystem after it successfully imports a statement for the first time and after sending ~Share~.
  • StatementDistributionMessage::NetworkBridgeUpdate
    • See next section.

Network bridge events

  • v1 compatibility
    • Messages for the v1 protocol are routed to the legacy statement distribution.
  • Statement
    • Notification of a signed statement.
    • Sent by a peer's Statement Distribution subsystem when circulating statements.
  • BackedCandidateManifest
    • Notification of a backed candidate being known by the sending node.
    • For the candidate being requested by the receiving node if needed.
    • Announcement.
    • Sent by a peer's Statement Distribution subsystem.
  • BackedCandidateKnown
    • Notification of a backed candidate being known by the sending node.
    • For informing a receiving node which already has the candidate.
    • Acknowledgement.
    • Sent by a peer's Statement Distribution subsystem.

Outgoing

  • NetworkBridgeTxMessage::SendValidationMessages
    • Sends a peer all pending messages / acknowledgements / statements for a relay parent, either through the cluster or the grid.
  • NetworkBridgeTxMessage::SendValidationMessage
    • Circulates a compact statement to all peers who need it, either through the cluster or the grid.
  • NetworkBridgeTxMessage::ReportPeer
    • Reports a peer (either good or bad).
  • CandidateBackingMessage::Statement
    • Note a validator's statement about a particular candidate.
  • ProspectiveParachainsMessage::GetHypotheticalFrontier
    • Gets the hypothetical frontier membership of candidates under active leaves' fragment trees.
  • NetworkBridgeTxMessage::SendRequests
    • Sends requests, initiating the request/response protocol.

Request/Response

We also have a request/response protocol because validators do not eagerly send each other heavy CommittedCandidateReceipt, but instead need to request these lazily.

Protocol

  1. Requesting Validator

    • Requests are queued up with RequestManager::get_or_insert.
      • Done as needed, when handling incoming manifests/statements.
    • RequestManager::dispatch_requests sends any queued-up requests.
      • Calls RequestManager::next_request to completion.
        • Creates the OutgoingRequest, saves the receiver in RequestManager::pending_responses.
      • Does nothing if we have more responses pending than the limit of parallel requests.
  2. Peer

    • Requests come in on a peer on the IncomingRequestReceiver.
      • Runs in a background responder task which feeds requests to answer_request through MuxedMessage.
      • This responder task has a limit on the number of parallel requests.
    • answer_request on the peer takes the request and sends a response.
      • Does this using the response sender on the request.
  3. Requesting Validator

    • receive_response on the original validator yields a response.
      • Response was sent on the request's response sender.
      • Uses RequestManager::await_incoming to await on pending responses in an unordered fashion.
      • Runs on the MuxedMessage receiver.
    • handle_response handles the response.

API

  • dispatch_requests
    • Dispatches pending requests for candidate data & statements.
  • answer_request
    • Answers an incoming request for a candidate.
    • Takes an incoming AttestedCandidateRequest.
  • receive_response
    • Wait on the next incoming response.
    • If there are no requests pending, this future never resolves.
    • Returns UnhandledResponse
  • handle_response
    • Handles an incoming response.
    • Takes UnhandledResponse

Manifests

A manifest is a message about a known backed candidate, along with a description of the statements backing it. It can be one of two kinds:

  • Full: Contains information about the candidate and should be sent to peers who may not have the candidate yet. This is also called an Announcement.
  • Acknowledgement: Omits information implicit in the candidate, and should be sent to peers which are guaranteed to have the candidate already.

Manifest Exchange

Manifest exchange is when a receiving node received a Full manifest and replied with an Acknowledgement. It indicates that both nodes know the candidate as valid and backed. This allows the nodes to send Statement messages directly to each other for any new statements.

Why? This limits the amount of statements we'd have to deal with w.r.t. candidates that don't really exist. Limiting out-of-group statement distribution between peers to only candidates that both peers agree are backed and exist ensures we only have to store statements about real candidates.

In practice, manifest exchange means that one of three things have happened:

  • They announced, we acknowledged.
  • We announced, they acknowledged.
  • We announced, they announced.

Concerning the last case, note that it is possible for two nodes to have each other in their sending set. Consider:

1 2
3 4

If validators 2 and 4 are in group B, then there is a path 2->1->3 and 4->3->1. Therefore, 1 and 3 might send each other manifests for the same candidate at the same time, without having seen the other's yet. This also counts as a manifest exchange, but is only allowed to occur in this way.

After the exchange is complete, we update pending statements. Pending statements are those we know locally that the remote node does not.

Alternative Paths Through The Topology

Nodes should send a BackedCandidateAcknowledgement(CandidateHash, StatementFilter) notification to any peer which has sent a manifest, and the candidate has been acquired by other means. This keeps alternative paths through the topology open, which allows nodes to receive additional statements that come later, but not after the candidate has been posted on-chain.

This is mostly about the limitation that the runtime has no way for block authors to post statements that come after the parablock is posted on-chain and ensure those validators still get rewarded. Technically, we only need enough statements to back the candidate and the manifest + request will provide that. But more statements might come shortly afterwards, and we want those to end up on-chain as well to ensure all validators in the group are rewarded.

For clarity, here is the full timeline:

  1. candidate seconded
  2. backable in cluster
  3. distributed along grid
  4. latecomers issue statements
  5. candidate posted on chain
  6. really latecomers issue statements

Cluster Module

The cluster module provides direct distribution of unbacked candidates within a group. By utilizing this initial phase of propagating only within clusters/groups, we bound the number of Seconded messages per validator per relay-parent, helping us prevent spam. Validators can try to circumvent this, but they would only consume a few KB of memory and it is trivially slashable on chain.

The cluster module determines whether to accept/reject messages from other validators in the same group. It keeps track of what we have sent to other validators in the group, and pending statements. For the full protocol, see "Protocol".

Grid Module

The grid module provides distribution of backed candidates and late statements outside the backing group. For the full protocol, see the "Protocol" section.

Grid Topology

For distributing outside our cluster (aka backing group) we use a 2D grid topology. This limits the amount of peers we send messages to, and handles view updates.

The basic operation of the grid topology is that:

  • A validator producing a message sends it to its row-neighbors and its column-neighbors.
  • A validator receiving a message originating from one of its row-neighbors sends it to its column-neighbors.
  • A validator receiving a message originating from one of its column-neighbors sends it to its row-neighbors.

This grid approach defines 2 unique paths for every validator to reach every other validator in at most 2 hops, providing redundancy.

Propagation follows these rules:

  • Each node has a receiving set and a sending set. These are different for each group. That is, if a node receives a candidate from group A, it checks if it is allowed to receive from that node for candidates from group A.
  • For groups that we are in, receive from nobody and send to our X/Y peers.
  • For groups that we are not part of:
    • We receive from any validator in the group we share a slice with and send to the corresponding X/Y slice in the other dimension.
    • For any validators we don't share a slice with, we receive from the nodes which share a slice with them.

Example

For size 11, the matrix would be:

0  1  2
3  4  5
6  7  8
9 10

e.g. for index 10, the neighbors would be 1, 4, 7, 9 -- these are the nodes we could directly communicate with (e.g. either send to or receive from).

Now, which of these neighbors can 10 receive from? Recall that the sending/receiving sets for 10 would be different for different groups. Here are some hypothetical scenarios:

  • Scenario 1: 9 belongs to group A but not 10. Here, 10 can directly receive candidates from group A from 9. 10 would propagate them to the nodes in {1, 4, 7} that are not in A.
  • Scenario 2: 6 is in group A instead of 9, and 7 is not in group A. 10 can receive group A messages from 7 or 9. 10 will try to relay these messages, but 7 and 9 together should have already propagated the message to all x/y peers of 10. If so, then 10 will just receive acknowledgements in reply rather than requests.
  • Scenario 3: 10 itself is in group A. 10 would not receive candidates from this group from any other nodes through the grid. It would itself send such candidates to all its neighbors that are not in A.

Seconding Limit

The seconding limit is a per-validator limit. Before asynchronous backing, we had a rule that every validator was only allowed to second one candidate per relay parent. With asynchronous backing, we have a 'maximum depth' which makes it possible to second multiple candidates per relay parent. The seconding limit is set to max depth + 1 to set an upper bound on candidates entering the system.

Candidates Module

The candidates module provides a tracker for all known candidates in the view, whether they are confirmed or not, and how peers have advertised the candidates. What is a confirmed candidate? It is a candidate for which we have the full receipt and the persisted validation data. This module gets confirmed candidates from two sources:

  • It can be that a validator fetched a collation directly from the collator and validated it.
  • The first time a validator gets an announcement for an unknown candidate, it will send a request for the candidate. Upon receiving a response and validating it (see UnhandledResponse::validate_response), it will mark the candidate as confirmed.

Requests Module

The requests module provides a manager for pending requests for candidate data, as well as pending responses. See "Request/Response Protocol" for a high-level description of the flow. See module-docs for full details.

Glossary

  • Acknowledgement: A partial manifest sent to a validator that already has the candidate to inform them that the sending node also knows the candidate. Concludes a manifest exchange.
  • Announcement: A full manifest indicating that a backed candidate is known by the sending node. Initiates a manifest exchange.
  • Attestation: See "Statement".
  • Backable vs. Backed:
    • Note that we sometimes use "backed" to refer to candidates that are "backable", but not yet backed on chain.
    • Backed should technically mean that the parablock candidate and its backing statements have been added to a relay chain block.
    • Backable is when the necessary backing statements have been acquired but those statements and the parablock candidate haven't been backed in a relay chain block yet.
  • Fragment tree: A parachain fragment not referenced by the relay-chain. It is a tree of prospective parachain blocks.
  • Manifest: A message about a known backed candidate, along with a description of the statements backing it. There are two kinds of manifest, Acknowledgement and Announcement. See "Manifests" section.
  • Peer: Another validator that a validator is connected to.
  • Request/response: A protocol used to lazily request and receive heavy candidate data when needed.
  • Reputation: Tracks reputation of peers. Applies annoyance cost and good behavior benefits.
  • Statement: Signed statements that can be made about parachain candidates.
    • Seconded: Proposal of a parachain candidate. Implicit validity vote.
    • Valid: States that a parachain candidate is valid.
  • Target: Target validator to send a statement to.
  • View: Current knowledge of the chain state.
    • Explicit view / immediate view
      • The view a peer has of the relay chain heads and highest finalized block.
    • Implicit view
      • Derived from the immediate view. Composed of active leaves and minimum relay-parents allowed for candidates of various parachains at those leaves.

Statement Distribution (Legacy)

This describes the legacy, backwards-compatible version of the Statement Distribution subsystem.

Note: All the V1 (legacy) code was extracted out to a legacy_v1 module of the statement-distribution crate, which doesn't alter any logic. V2 (new protocol) peers also run legacy_v1 and communicate with V1 peers using V1 messages and with V2 peers using V2 messages. Once the runtime upgrade goes through on all networks, this legacy_v1 code will no longer be triggered and will be vestigial and can be removed.

Overview

The Statement Distribution Subsystem is responsible for distributing statements about seconded candidates between validators.

Protocol

PeerSet: Validation

Input:

  • NetworkBridgeUpdate(update)
  • StatementDistributionMessage

Output:

  • NetworkBridge::SendMessage(PeerId, message)
  • NetworkBridge::SendRequests(StatementFetchingV1)
  • NetworkBridge::ReportPeer(PeerId, cost_or_benefit)

Functionality

Implemented as a gossip protocol. Handles updates to our view and peers' views. Neighbor packets are used to inform peers which chain heads we are interested in data for.

The Statement Distribution Subsystem is responsible for distributing signed statements that we have generated and for forwarding statements generated by other validators. It also detects a variety of Validator misbehaviors for reporting to the Provisioner Subsystem. During the Backing stage of the inclusion pipeline, Statement Distribution is the main point of contact with peer nodes. On receiving a signed statement from a peer in the same backing group, assuming the peer receipt state machine is in an appropriate state, it sends the Candidate Receipt to the Candidate Backing subsystem to handle the validator's statement. On receiving StatementDistributionMessage::Share we make sure to send messages to our backing group in addition to random other peers, to ensure a fast backing process and getting all statements quickly for distribution.

This subsystem tracks equivocating validators and stops accepting information from them. It establishes a data-dependency order:

  • In order to receive a Seconded message we have the corresponding chain head in our view
  • In order to receive a Valid message we must have received the corresponding Seconded message.

And respect this data-dependency order from our peers by respecting their views. This subsystem is responsible for checking message signatures.

The Statement Distribution subsystem sends statements to peer nodes.

Peer Receipt State Machine

There is a very simple state machine which governs which messages we are willing to receive from peers. Not depicted in the state machine: on initial receipt of any SignedFullStatement, validate that the provided signature does in fact sign the included data. Note that each individual parablock candidate gets its own instance of this state machine; it is perfectly legal to receive a Valid(X) before a Seconded(Y), as long as a Seconded(X) has been received.

A: Initial State. Receive SignedFullStatement(Statement::Second): extract Statement, forward to Candidate Backing, proceed to B. Receive any other SignedFullStatement variant: drop it.

B: Receive any SignedFullStatement: check signature and determine whether the statement is new to us. if new, forward to Candidate Backing and circulate to other peers. Receive OverseerMessage::StopWork: proceed to C.

C: Receive any message for this block: drop it.

For large statements (see below), we also keep track of the total received large statements per peer and have a hard limit on that number for flood protection. This is necessary as in the current code we only forward statements once we have all the data, therefore flood protection for large statement is a bit more subtle. This will become an obsolete problem once off chain code upgrades are implemented.

Peer Knowledge Tracking

The peer receipt state machine implies that for parsimony of network resources, we should model the knowledge of our peers, and help them out. For example, let's consider a case with peers A, B, and C, validators X and Y, and candidate M. A sends us a Statement::Second(M) signed by X. We've double-checked it, and it's valid. While we're checking it, we receive a copy of X's Statement::Second(M) from B, along with a Statement::Valid(M) signed by Y.

Our response to A is just the Statement::Valid(M) signed by Y. However, we haven't heard anything about this from C. Therefore, we send it everything we have: first a copy of X's Statement::Second, then Y's Statement::Valid.

This system implies a certain level of duplication of messages--we received X's Statement::Second from both our peers, and C may experience the same--but it minimizes the degree to which messages are simply dropped.

And respect this data-dependency order from our peers. This subsystem is responsible for checking message signatures.

No jobs. We follow view changes from the NetworkBridge, which in turn is updated by the overseer.

Equivocations and Flood Protection

An equivocation is a double-vote by a validator. The Candidate Backing Subsystem is better-suited than this one to detect equivocations as it adds votes to quorum trackers.

At this level, we are primarily concerned about flood-protection, and to some extent, detecting equivocations is a part of that. In particular, we are interested in detecting equivocations of Seconded statements. Since every other statement is dependent on Seconded statements, ensuring that we only ever hold a bounded number of Seconded statements is sufficient for flood-protection.

The simple approach is to say that we only receive up to two Seconded statements per validator per chain head. However, the marginal cost of equivocation, conditional on having already equivocated, is close to 0, since a single double-vote offence is counted as all double-vote offences for a particular chain-head. Even if it were not, there is some amount of equivocations that can be done such that the marginal cost of issuing further equivocations is close to 0, as there would be an amount of equivocations necessary to be completely and totally obliterated by the slashing algorithm. We fear the validator with nothing left to lose.

With that in mind, this simple approach has a caveat worth digging deeper into.

First: We may be aware of two equivocated Seconded statements issued by a validator. A totally honest peer of ours can also be aware of one or two different Seconded statements issued by the same validator. And yet another peer may be aware of one or two more Seconded statements. And so on. This interacts badly with pre-emptive sending logic. Upon sending a Seconded statement to a peer, we will want to pre-emptively follow up with all statements relative to that candidate. Waiting for acknowledgment introduces latency at every hop, so that is best avoided. What can happen is that upon receipt of the Seconded statement, the peer will discard it as it falls beyond the bound of 2 that it is allowed to store. It cannot store anything in memory about discarded candidates as that would introduce a DoS vector. Then, the peer would receive from us all of the statements pertaining to that candidate, which, from its perspective, would be undesired - they are data-dependent on the Seconded statement we sent them, but they have erased all record of that from their memory. Upon receiving a potential flood of undesired statements, this 100% honest peer may choose to disconnect from us. In this way, an adversary may be able to partition the network with careful distribution of equivocated Seconded statements.

The fix is to track, per-peer, the hashes of up to 4 candidates per validator (per relay-parent) that the peer is aware of. It is 4 because we may send them 2 and they may send us 2 different ones. We track the data that they are aware of as the union of things we have sent them and things they have sent us. If we receive a 1st or 2nd Seconded statement from a peer, we note it in the peer's known candidates even if we do disregard the data locally. And then, upon receipt of any data dependent on that statement, we do not reduce that peer's standing in our eyes, as the data was not undesired.

There is another caveat to the fix: we don't want to allow the peer to flood us because it has set things up in a way that it knows we will drop all of its traffic. We also track how many statements we have received per peer, per candidate, and per chain-head. This is any statement concerning a particular candidate: Seconded, Valid, or Invalid. If we ever receive a statement from a peer which would push any of these counters beyond twice the amount of validators at the chain-head, we begin to lower the peer's standing and eventually disconnect. This bound is a massive overestimate and could be reduced to twice the number of validators in the corresponding validator group. It is worth noting that the goal at the time of writing is to ensure any finite bound on the amount of stored data, as any equivocation results in a large slash.

Large statements

Seconded statements can become quite large on parachain runtime upgrades for example. For this reason, there exists a LargeStatement constructor for the StatementDistributionMessage wire message, which only contains light metadata of a statement. The actual candidate data is not included. This message type is used whenever a message is deemed large. The receiver of such a message needs to request the actual payload via request/response by means of a StatementFetchingV1 request.

This is necessary as distribution of a large payload (mega bytes) via gossip would make the network collapse and timely distribution of statements would no longer be possible. By using request/response it is ensured that each peer only transferes large data once. We only take good care to detect an overloaded peer early and immediately move on to a different peer for fetching the data. This mechanism should result in a good load distribution and therefore a rather optimal distribution path.

With these optimizations, distribution of payloads in the size of up to 3 to 4 MB should work with Kusama validator specifications. For scaling up even more, runtime upgrades and message passing should be done off chain at some point.

Flood protection considerations: For making DoS attacks slightly harder on this subsystem, nodes will only respond to large statement requests, when they previously notified that peer via gossip about that statement. So, it is not possible to DoS nodes at scale, by requesting candidate data over and over again.

Availability Subsystems

The availability subsystems are responsible for ensuring that Proofs of Validity of backed candidates are widely available within the validator set, without requiring every node to retain a full copy. They accomplish this by broadly distributing erasure-coded chunks of the PoV, keeping track of which validator has which chunk by means of signed bitfields. They are also responsible for reassembling a complete PoV when required, e.g. when an approval checker needs to validate a parachain block.

Availability Distribution

This subsystem is responsible for distribution availability data to peers. Availability data are chunks, PoVs and AvailableData (which is PoV + PersistedValidationData). It does so via request response protocols.

In particular this subsystem is responsible for:

  • Respond to network requests requesting availability data by querying the Availability Store.
  • Request chunks from backing validators to put them in the local Availability Store whenever we find an occupied core on any fresh leaf, this is to ensure availability by at least 2/3+ of all validators, this happens after a candidate is backed.
  • Fetch PoV from validators, when requested via FetchPoV message from backing (pov_requester module).

The backing subsystem is responsible of making available data available in the local Availability Store upon validation. This subsystem will serve any network requests by querying that store.

Protocol

This subsystem does not handle any peer set messages, but the pov_requester does connect to validators of the same backing group on the validation peer set, to ensure fast propagation of statements between those validators and for ensuring already established connections for requesting PoVs. Other than that this subsystem drives request/response protocols.

Input:

  • OverseerSignal::ActiveLeaves(ActiveLeavesUpdate)
  • AvailabilityDistributionMessage{msg: ChunkFetchingRequest}
  • AvailabilityDistributionMessage{msg: PoVFetchingRequest}
  • AvailabilityDistributionMessage{msg: FetchPoV}

Output:

  • NetworkBridgeMessage::SendRequests(Requests, IfDisconnected::TryConnect)
  • AvailabilityStore::QueryChunk(candidate_hash, index, response_channel)
  • AvailabilityStore::StoreChunk(candidate_hash, chunk)
  • AvailabilityStore::QueryAvailableData(candidate_hash, response_channel)
  • RuntimeApiRequest::SessionIndexForChild
  • RuntimeApiRequest::SessionInfo
  • RuntimeApiRequest::AvailabilityCores

Functionality

PoV Requester

The PoV requester in the pov_requester module takes care of staying connected to validators of the current backing group of this very validator on the Validation peer set and it will handle FetchPoV requests by issuing network requests to those validators. It will check the hash of the received PoV, but will not do any further validation. That needs to be done by the original FetchPoV sender (backing subsystem).

Chunk Requester

After a candidate is backed, the availability of the PoV block must be confirmed by 2/3+ of all validators. The chunk requester is responsible of making that availability a reality.

It does that by querying checking occupied cores for all active leaves. For each occupied core it will spawn a task fetching the erasure chunk which has the ValidatorIndex of the node. For this an ChunkFetchingRequest is issued, via substrate's generic request/response protocol.

The spawned task will start trying to fetch the chunk from validators in responsible group of the occupied core, in a random order. For ensuring that we use already open TCP connections wherever possible, the requester maintains a cache and preserves that random order for the entire session.

Note however that, because not all validators in a group have to be actual backers, not all of them are required to have the needed chunk. This in turn could lead to low throughput, as we have to wait for fetches to fail, before reaching a validator finally having our chunk. We do rank back validators not delivering our chunk, but as backers could vary from block to block on a perfectly legitimate basis, this is still not ideal. See issues 2509 and 2512 for more information.

The current implementation also only fetches chunks for occupied cores in blocks in active leaves. This means though, if active leaves skips a block or we are particularly slow in fetching our chunk, we might not fetch our chunk if availability reached 2/3 fast enough (slot becomes free). This is not desirable as we would like as many validators as possible to have their chunk. See this issue for more details.

Serving

On the other side the subsystem will listen for incoming ChunkFetchingRequests and PoVFetchingRequests from the network bridge and will respond to queries, by looking the requested chunks and PoVs up in the availability store, this happens in the responder module.

We rely on the backing subsystem to make available data available locally in the Availability Store after it has validated it.

Availability Recovery

This subsystem is the inverse of the Availability Distribution subsystem: validators will serve the availability chunks kept in the availability store to nodes who connect to them. And the subsystem will also implement the other side: the logic for nodes to connect to validators, request availability pieces, and reconstruct the AvailableData.

This version of the availability recovery subsystem is based off of direct connections to validators. In order to recover any given AvailableData, we must recover at least f + 1 pieces from validators of the session. Thus, we will connect to and query randomly chosen validators until we have received f + 1 pieces.

Protocol

PeerSet: Validation

Input:

  • NetworkBridgeUpdate(update)
  • AvailabilityRecoveryMessage::RecoverAvailableData(candidate, session, backing_group, response)

Output:

  • NetworkBridge::SendValidationMessage
  • NetworkBridge::ReportPeer
  • AvailabilityStore::QueryChunk

Functionality

We hold a state which tracks the currently ongoing recovery tasks, as well as which request IDs correspond to which task. A recovery task is a structure encapsulating all recovery tasks with the network necessary to recover the available data in respect to one candidate.

#![allow(unused)]
fn main() {
struct State {
    /// Each recovery is implemented as an independent async task, and the handles only supply information about the result.
    ongoing_recoveries: FuturesUnordered<RecoveryHandle>,
    /// A recent block hash for which state should be available.
    live_block_hash: Hash,
    // An LRU cache of recently recovered data.
    availability_lru: LruCache<CandidateHash, Result<AvailableData, RecoveryError>>,
}

/// This is a future, which concludes either when a response is received from the recovery tasks,
/// or all the `awaiting` channels have closed.
struct RecoveryHandle {
    candidate_hash: CandidateHash,
    interaction_response: RemoteHandle<Concluded>,
    awaiting: Vec<ResponseChannel<Result<AvailableData, RecoveryError>>>,
}

struct Unavailable;
struct Concluded(CandidateHash, Result<AvailableData, RecoveryError>);

struct RecoveryTaskParams {
    validator_authority_keys: Vec<AuthorityId>,
    validators: Vec<ValidatorId>,
    // The number of pieces needed.
    threshold: usize,
    candidate_hash: Hash,
    erasure_root: Hash,
}

enum RecoveryTask {
    RequestFromBackers {
        // a random shuffling of the validators from the backing group which indicates the order
        // in which we connect to them and request the chunk.
        shuffled_backers: Vec<ValidatorIndex>,
    }
    RequestChunksFromValidators {
        // a random shuffling of the validators which indicates the order in which we connect to the validators and
        // request the chunk from them.
        shuffling: Vec<ValidatorIndex>,
        received_chunks: Map<ValidatorIndex, ErasureChunk>,
        requesting_chunks: FuturesUnordered<Receiver<ErasureChunkRequestResponse>>,
    }
}

struct RecoveryTask {
    to_subsystems: SubsystemSender,
    params: RecoveryTaskParams,
    source: Source,
}
}

Signal Handling

On ActiveLeavesUpdate, if activated is non-empty, set state.live_block_hash to the first block in Activated.

Ignore BlockFinalized signals.

On Conclude, shut down the subsystem.

AvailabilityRecoveryMessage::RecoverAvailableData(receipt, session, Option<backing_group_index>, response)

  1. Check the availability_lru for the candidate and return the data if so.
  2. Check if there is already an recovery handle for the request. If so, add the response handle to it.
  3. Otherwise, load the session info for the given session under the state of live_block_hash, and initiate a recovery task with launch_recovery_task. Add a recovery handle to the state and add the response channel to it.
  4. If the session info is not available, return RecoveryError::Unavailable on the response channel.

Recovery logic

launch_recovery_task(session_index, session_info, candidate_receipt, candidate_hash, Option<backing_group_index>)

  1. Compute the threshold from the session info. It should be f + 1, where n = 3f + k, where k in {1, 2, 3}, and n is the number of validators.
  2. Set the various fields of RecoveryParams based on the validator lists in session_info and information about the candidate.
  3. If the backing_group_index is Some, start in the RequestFromBackers phase with a shuffling of the backing group validator indices and a None requesting value.
  4. Otherwise, start in the RequestChunksFromValidators source with received_chunks,requesting_chunks, and next_shuffling all empty.
  5. Set the to_subsystems sender to be equal to a clone of the SubsystemContext's sender.
  6. Initialize received_chunks to an empty set, as well as requesting_chunks.

Launch the source as a background task running run(recovery_task).

run(recovery_task) -> Result<AvailableData, RecoeryError>

#![allow(unused)]
fn main() {
// How many parallel requests to have going at once.
const N_PARALLEL: usize = 50;
}
  • Request AvailabilityStoreMessage::QueryAvailableData. If it exists, return that.

  • If the task contains RequestFromBackers

    • Loop:
      • If the requesting_pov is Some, poll for updates on it. If it concludes, set requesting_pov to None.
      • If the requesting_pov is None, take the next backer off the shuffled_backers.
        • If the backer is Some, issue a NetworkBridgeMessage::Requests with a network request for the AvailableData and wait for the response.
        • If it concludes with a None result, return to beginning.
        • If it concludes with available data, attempt a re-encoding.
          • If it has the correct erasure-root, break and issue a Ok(available_data).
          • If it has an incorrect erasure-root, return to beginning.
        • Send the result to each member of awaiting.
        • If the backer is None, set the source to RequestChunksFromValidators with a random shuffling of validators and empty received_chunks, and requesting_chunks and break the loop.
  • If the task contains RequestChunksFromValidators:

    • Request AvailabilityStoreMessage::QueryAllChunks. For each chunk that exists, add it to received_chunks and remote the validator from shuffling.
    • Loop:
      • If received_chunks + requesting_chunks + shuffling lengths are less than the threshold, break and return Err(Unavailable).
      • Poll for new updates from requesting_chunks. Check merkle proofs of any received chunks. If the request simply fails due to network issues, insert into the front of shuffling to be retried.
      • If received_chunks has more than threshold entries, attempt to recover the data.
        • If that fails, return Err(RecoveryError::Invalid)
        • If correct:
          • If re-encoding produces an incorrect erasure-root, break and issue a Err(RecoveryError::Invalid).
          • break and issue Ok(available_data)
      • Send the result to each member of awaiting.
      • While there are fewer than N_PARALLEL entries in requesting_chunks,
        • Pop the next item from shuffling. If it's empty and requesting_chunks is empty, return Err(RecoveryError::Unavailable).
        • Issue a NetworkBridgeMessage::Requests and wait for the response in requesting_chunks.

Bitfield Distribution

Validators vote on the availability of a backed candidate by issuing signed bitfields, where each bit corresponds to a single candidate. These bitfields can be used to compactly determine which backed candidates are available or not based on a 2/3+ quorum.

Protocol

PeerSet: Validation

Input: BitfieldDistributionMessage which are gossiped to all peers, no matter if validator or not.

Output:

  • NetworkBridge::SendValidationMessage([PeerId], message) gossip a verified incoming bitfield on to interested subsystems within this validator node.
  • NetworkBridge::ReportPeer(PeerId, cost_or_benefit) improve or penalize the reputation of peers based on the messages that are received relative to the current view.
  • ProvisionerMessage::ProvisionableData(ProvisionableData::Bitfield(relay_parent, SignedAvailabilityBitfield)) pass on the bitfield to the other submodules via the overseer.

Functionality

This is implemented as a gossip system.

It is necessary to track peer connection, view change, and disconnection events, in order to maintain an index of which peers are interested in which relay parent bitfields.

Before gossiping incoming bitfields, they must be checked to be signed by one of the validators of the validator set relevant to the current relay parent. Only accept bitfields relevant to our current view and only distribute bitfields to other peers when relevant to their most recent view. Accept and distribute only one bitfield per validator.

When receiving a bitfield either from the network or from a DistributeBitfield message, forward it along to the block authorship (provisioning) subsystem for potential inclusion in a block.

Peers connecting after a set of valid bitfield gossip messages was received, those messages must be cached and sent upon connection of new peers or re-connecting peers.

Bitfield Signing

Validators vote on the availability of a backed candidate by issuing signed bitfields, where each bit corresponds to a single candidate. These bitfields can be used to compactly determine which backed candidates are available or not based on a 2/3+ quorum.

Protocol

Input:

There is no dedicated input mechanism for bitfield signing. Instead, Bitfield Signing produces a bitfield representing the current state of availability on StartWork.

Output:

  • BitfieldDistribution::DistributeBitfield: distribute a locally signed bitfield
  • AvailabilityStore::QueryChunk(CandidateHash, validator_index, response_channel)

Functionality

Upon receipt of an ActiveLeavesUpdate, launch bitfield signing job for each activated head referring to a fresh leaf. Stop the job for each deactivated head.

Bitfield Signing Job

Localized to a specific relay-parent r If not running as a validator, do nothing.

  • For each fresh leaf, begin by waiting a fixed period of time so availability distribution has the chance to make candidates available.
  • Determine our validator index i, the set of backed candidates pending availability in r, and which bit of the bitfield each corresponds to.
  • Start with an empty bitfield. For each bit in the bitfield, if there is a candidate pending availability, query the Availability Store for whether we have the availability chunk for our validator index. The OccupiedCore struct contains the candidate hash so the full candidate does not need to be fetched from runtime.
  • For all chunks we have, set the corresponding bit in the bitfield.
  • Sign the bitfield and dispatch a BitfieldDistribution::DistributeBitfield message.

Approval Subsystems

The approval subsystems implement the node-side of the Approval Protocol.

We make a divide between the assignment/voting logic and the distribution logic that distributes assignment certifications and approval votes. The logic in the assignment and voting also informs the GRANDPA voting rule on how to vote.

These subsystems are intended to flag issues and begin participating in live disputes. Dispute subsystems also track all observed votes (backing, approval, and dispute-specific) by all validators on all candidates.

Approval Voting

Reading the section on the approval protocol will likely be necessary to understand the aims of this subsystem.

Approval votes are split into two parts: Assignments and Approvals. Validators first broadcast their assignment to indicate intent to check a candidate. Upon successfully checking, they broadcast an approval vote. If a validator doesn't broadcast their approval vote shortly after issuing an assignment, this is an indication that they are being prevented from recovering or validating the block data and that more validators should self-select to check the candidate. This is known as a "no-show".

The core of this subsystem is a Tick-based timer loop, where Ticks are 500ms. We also reason about time in terms of DelayTranches, which measure the number of ticks elapsed since a block was produced. We track metadata for all un-finalized but included candidates. We compute our local assignments to check each candidate, as well as which DelayTranche those assignments may be minimally triggered at. As the same candidate may appear in more than one block, we must produce our potential assignments for each (Block, Candidate) pair. The timing loop is based on waiting for assignments to become no-shows or waiting to broadcast and begin our own assignment to check.

Another main component of this subsystem is the logic for determining when a (Block, Candidate) pair has been approved and when to broadcast and trigger our own assignment. Once a (Block, Candidate) pair has been approved, we mark a corresponding bit in the BlockEntry that indicates the candidate has been approved under the block. When we trigger our own assignment, we broadcast it via Approval Distribution, begin fetching the data from Availability Recovery, and then pass it through to the Candidate Validation. Once these steps are successful, we issue our approval vote. If any of these steps fail, we don't issue any vote and will "no-show" from the perspective of other validators in addition a dispute is raised via the dispute-coordinator, by sending IssueLocalStatement.

Where this all fits into Polkadot is via block finality. Our goal is to not finalize any block containing a candidate that is not approved. We provide a hook for a custom GRANDPA voting rule - GRANDPA makes requests of the form (target, minimum) consisting of a target block (i.e. longest chain) that it would like to finalize, and a minimum block which, due to the rules of GRANDPA, must be voted on. The minimum is typically the last finalized block, but may be beyond it, in the case of having a last-round-estimate beyond the last finalized. Thus, our goal is to inform GRANDPA of some block between target and minimum which we believe can be finalized safely. We do this by iterating backwards from the target to the minimum and finding the longest continuous chain from minimum where all candidates included by those blocks have been approved.

Protocol

Input:

  • ApprovalVotingMessage::CheckAndImportAssignment
  • ApprovalVotingMessage::CheckAndImportApproval
  • ApprovalVotingMessage::ApprovedAncestor

Output:

  • ApprovalDistributionMessage::DistributeAssignment
  • ApprovalDistributionMessage::DistributeApproval
  • RuntimeApiMessage::Request
  • ChainApiMessage
  • AvailabilityRecoveryMessage::Recover
  • CandidateExecutionMessage::ValidateFromExhaustive

Functionality

The approval voting subsystem is responsible for casting votes and determining approval of candidates and as a result, blocks.

This subsystem wraps a database which is used to store metadata about unfinalized blocks and the candidates within them. Candidates may appear in multiple blocks, and assignment criteria are chosen differently based on the hash of the block they appear in.

Database Schema

The database schema is designed with the following goals in mind:

  1. To provide an easy index from unfinalized blocks to candidates
  2. To provide a lookup from candidate hash to approval status
  3. To be easy to clear on start-up. What has happened while we were offline is unimportant.
  4. To be fast to clear entries outdated by finality

Structs:

#![allow(unused)]
fn main() {
struct TrancheEntry {
    tranche: DelayTranche,
    // assigned validators who have not yet approved, and the instant we received
    // their assignment.
    assignments: Vec<(ValidatorIndex, Tick)>,
}

struct OurAssignment {
  cert: AssignmentCert,
  tranche: DelayTranche,
  validator_index: ValidatorIndex,
  triggered: bool,
}

struct ApprovalEntry {
    tranches: Vec<TrancheEntry>, // sorted ascending by tranche number.
    backing_group: GroupIndex,
    our_assignment: Option<OurAssignment>,
    our_approval_sig: Option<ValidatorSignature>,
    assignments: Bitfield, // n_validators bits
    approved: bool,
}

struct CandidateEntry {
    candidate: CandidateReceipt,
    session: SessionIndex,
    // Assignments are based on blocks, so we need to track assignments separately
    // based on the block we are looking at.
    block_assignments: HashMap<Hash, ApprovalEntry>,
    approvals: Bitfield, // n_validators bits
}

struct BlockEntry {
    block_hash: Hash,
    session: SessionIndex,
    slot: Slot,
    // random bytes derived from the VRF submitted within the block by the block
    // author as a credential and used as input to approval assignment criteria.
    relay_vrf_story: [u8; 32],
    // The candidates included as-of this block and the index of the core they are
    // leaving. Sorted ascending by core index.
    candidates: Vec<(CoreIndex, Hash)>,
    // A bitfield where the i'th bit corresponds to the i'th candidate in `candidates`.
    // The i'th bit is `true` iff the candidate has been approved in the context of
    // this block. The block can be considered approved has all bits set to 1
    approved_bitfield: Bitfield,
    children: Vec<Hash>,
}

// slot_duration * 2 + DelayTranche gives the number of delay tranches since the
// unix epoch.
type Tick = u64;

struct StoredBlockRange(BlockNumber, BlockNumber);
}

In the schema, we map

"StoredBlocks" => StoredBlockRange
BlockNumber => Vec<BlockHash>
BlockHash => BlockEntry
CandidateHash => CandidateEntry

Logic

#![allow(unused)]
fn main() {
const APPROVAL_SESSIONS: SessionIndex = 6;

// The minimum amount of ticks that an assignment must have been known for.
const APPROVAL_DELAY: Tick = 2;
}

In-memory state:

#![allow(unused)]
fn main() {
struct ApprovalVoteRequest {
  validator_index: ValidatorIndex,
  block_hash: Hash,
  candidate_index: CandidateIndex,
}

// Requests that background work (approval voting tasks) may need to make of the main subsystem
// task.
enum BackgroundRequest {
  ApprovalVote(ApprovalVoteRequest),
  // .. others, unspecified as per implementation.
}

// This is the general state of the subsystem. The actual implementation may split this
// into further pieces.
struct State {
    earliest_session: SessionIndex,
    session_info: Vec<SessionInfo>,
    babe_epoch: Option<BabeEpoch>, // information about a cached BABE epoch.
    keystore: Keystore,

    // A scheduler which keeps at most one wakeup per hash, candidate hash pair and
    // maps such pairs to `Tick`s.
    wakeups: Wakeups,

    // These are connected to each other.
    background_tx: mpsc::Sender<BackgroundRequest>,
    background_rx: mpsc::Receiver<BackgroundRequest>,
}
}

This guide section makes no explicit references to writes to or reads from disk. Instead, it handles them implicitly, with the understanding that updates to block, candidate, and approval entries are persisted to disk.

SessionInfo

On start-up, we clear everything currently stored by the database. This is done by loading the StoredBlockRange, iterating through each block number, iterating through each block hash, and iterating through each candidate referenced by each block. Although this is O(o*n*p), we don't expect to have more than a few unfinalized blocks at any time and in extreme cases, a few thousand. The clearing operation should be relatively fast as a result.

Main loop:

  • Each iteration, select over all of
    • The next Tick in wakeups: trigger wakeup_process for each (Hash, Hash) pair scheduled under the Tick and then remove all entries under the Tick.
    • The next message from the overseer: handle the message as described in the Incoming Messages section
    • The next approval vote request from background_rx

Incoming Messages

OverseerSignal::BlockFinalized

On receiving an OverseerSignal::BlockFinalized(h), we fetch the block number b of that block from the ChainApi subsystem. We update our StoredBlockRange to begin at b+1. Additionally, we remove all block entries and candidates referenced by them up to and including b. Lastly, we prune out all descendants of h transitively: when we remove a BlockEntry with number b that is not equal to h, we recursively delete all the BlockEntrys referenced as children. We remove the block_assignments entry for the block hash and if block_assignments is now empty, remove the CandidateEntry. We also update each of the BlockNumber -> Vec<Hash> keys in the database to reflect the blocks at that height, clearing if empty.

OverseerSignal::ActiveLeavesUpdate

On receiving an OverseerSignal::ActiveLeavesUpdate(update):

  • We determine the set of new blocks that were not in our previous view. This is done by querying the ancestry of all new items in the view and contrasting against the stored BlockNumbers. Typically, there will be only one new block. We fetch the headers and information on these blocks from the ChainApi subsystem. Stale leaves in the update can be ignored.
  • We update the StoredBlockRange and the BlockNumber maps.
  • We use the RuntimeApiSubsystem to determine information about these blocks. It is generally safe to assume that runtime state is available for recent, unfinalized blocks. In the case that it isn't, it means that we are catching up to the head of the chain and needn't worry about assignments to those blocks anyway, as the security assumption of the protocol tolerates nodes being temporarily offline or out-of-date.
    • We fetch the set of candidates included by each block by dispatching a RuntimeApiRequest::CandidateEvents and checking the CandidateIncluded events.
    • We fetch the session of the block by dispatching a session_index_for_child request with the parent-hash of the block.
    • If the session index - APPROVAL_SESSIONS > state.earliest_session, then bump state.earliest_sessions to that amount and prune earlier sessions.
    • If the session isn't in our state.session_info, load the session info for it and for all sessions since the earliest-session, including the earliest-session, if that is missing. And it can be, just after pruning, if we've done a big jump forward, as is the case when we've just finished chain synchronization.
    • If any of the runtime API calls fail, we just warn and skip the block.
  • We use the RuntimeApiSubsystem to determine the set of candidates included in these blocks and use BABE logic to determine the slot number and VRF of the blocks.
  • We also note how late we appear to have received the block. We create a BlockEntry for each block and a CandidateEntry for each candidate obtained from CandidateIncluded events after making a RuntimeApiRequest::CandidateEvents request.
  • For each candidate, if the amount of needed approvals is more than the validators remaining after the backing group of the candidate is subtracted, then the candidate is insta-approved as approval would be impossible otherwise. If all candidates in the block are insta-approved, or there are no candidates in the block, then the block is insta-approved. If the block is insta-approved, a ChainSelectionMessage::Approved should be sent for the block.
  • Ensure that the CandidateEntry contains a block_assignments entry for the block, with the correct backing group set.
  • If a validator in this session, compute and assign our_assignment for the block_assignments
    • Only if not a member of the backing group.
    • Run RelayVRFModulo and RelayVRFDelay according to the the approvals protocol section. Ensure that the assigned core derived from the output is covered by the auxiliary signature aggregated in the VRFPRoof.
  • Handle Wakeup for each new candidate in each new block - this will automatically broadcast a 0-tranche assignment, kick off approval work, and schedule the next delay.
  • Dispatch an ApprovalDistributionMessage::NewBlocks with the meta information filled out for each new block.

ApprovalVotingMessage::CheckAndImportAssignment

On receiving a ApprovalVotingMessage::CheckAndImportAssignment message, we check the assignment cert against the block entry. The cert itself contains information necessary to determine the candidate that is being assigned-to. In detail:

  • Load the BlockEntry for the relay-parent referenced by the message. If there is none, return AssignmentCheckResult::Bad.
  • Fetch the SessionInfo for the session of the block
  • Determine the assignment key of the validator based on that.
  • Determine the claimed core index by looking up the candidate with given index in block_entry.candidates. Return AssignmentCheckResult::Bad if missing.
  • Check the assignment cert
    • If the cert kind is RelayVRFModulo, then the certificate is valid as long as sample < session_info.relay_vrf_samples and the VRF is valid for the validator's key with the input block_entry.relay_vrf_story ++ sample.encode() as described with the approvals protocol section. We set core_index = vrf.make_bytes().to_u32() % session_info.n_cores. If the BlockEntry causes inclusion of a candidate at core_index, then this is a valid assignment for the candidate at core_index and has delay tranche 0. Otherwise, it can be ignored.
    • If the cert kind is RelayVRFDelay, then we check if the VRF is valid for the validator's key with the input block_entry.relay_vrf_story ++ cert.core_index.encode() as described in the approvals protocol section. The cert can be ignored if the block did not cause inclusion of a candidate on that core index. Otherwise, this is a valid assignment for the included candidate. The delay tranche for the assignment is determined by reducing (vrf.make_bytes().to_u64() % (session_info.n_delay_tranches + session_info.zeroth_delay_tranche_width)).saturating_sub(session_info.zeroth_delay_tranche_width).
    • We also check that the core index derived by the output is covered by the VRFProof by means of an auxiliary signature.
    • If the delay tranche is too far in the future, return AssignmentCheckResult::TooFarInFuture.
  • Import the assignment.
    • Load the candidate in question and access the approval_entry for the block hash the cert references.
    • Ignore if we already observe the validator as having been assigned.
    • Ensure the validator index is not part of the backing group for the candidate.
    • Ensure the validator index is not present in the approval entry already.
    • Create a tranche entry for the delay tranche in the approval entry and note the assignment within it.
    • Note the candidate index within the approval entry.
  • Schedule a wakeup for this block, candidate pair.
  • return the appropriate AssignmentCheckResult on the response channel.

ApprovalVotingMessage::CheckAndImportApproval

On receiving a CheckAndImportApproval(indirect_approval_vote, response_channel) message:

  • Fetch the BlockEntry from the indirect approval vote's block_hash. If none, return ApprovalCheckResult::Bad.
  • Fetch the CandidateEntry from the indirect approval vote's candidate_index. If the block did not trigger inclusion of enough candidates, return ApprovalCheckResult::Bad.
  • Construct a SignedApprovalVote using the candidate hash and check against the validator's approval key, based on the session info of the block. If invalid or no such validator, return ApprovalCheckResult::Bad.
  • Send ApprovalCheckResult::Accepted
  • Import the checked approval vote

ApprovalVotingMessage::ApprovedAncestor

On receiving an ApprovedAncestor(Hash, BlockNumber, response_channel):

  • Iterate over the ancestry of the hash all the way back to block number given, starting from the provided block hash. Load the CandidateHashes from each block entry.
  • Keep track of an all_approved_max: Option<(Hash, BlockNumber, Vec<(Hash, Vec<CandidateHash>))>.
  • For each block hash encountered, load the BlockEntry associated. If any are not found, return None on the response channel and conclude.
  • If the block entry's approval_bitfield has all bits set to 1 and all_approved_max == None, set all_approved_max = Some((current_hash, current_number)).
  • If the block entry's approval_bitfield has any 0 bits, set all_approved_max = None.
  • If all_approved_max is Some, push the current block hash and candidate hashes onto the list of blocks and candidates all_approved_max.
  • After iterating all ancestry, return all_approved_max.

Updates and Auxiliary Logic

Import Checked Approval

  • Import an approval vote which we can assume to have passed signature checks and correspond to an imported assignment.
  • Requires (BlockEntry, CandidateEntry, ValidatorIndex)
  • Set the corresponding bit of the approvals bitfield in the CandidateEntry to 1. If already 1, return.
  • Checks the approval state of a candidate under a specific block, and updates the block and candidate entries accordingly.
  • Checks the ApprovalEntry for the block.
  • If the approval vote originates locally, set the our_approval_sig in the candidate entry.

Handling Wakeup

  • Handle a previously-scheduled wakeup of a candidate under a specific block.
  • Requires (relay_block, candidate_hash)
  • Load the BlockEntry and CandidateEntry from disk. If either is not present, this may have lost a race with finality and can be ignored. Also load the ApprovalEntry for the block and candidate.
  • determine the RequiredTranches of the candidate.
  • Determine if we should trigger our assignment.
    • If we've already triggered or OurAssignment is None, we do not trigger.
    • If we have RequiredTranches::All, then we trigger if the candidate is not approved. We have no next wakeup as we assume that other validators are doing the same and we will be implicitly woken up by handling new votes.
    • If we have RequiredTranches::Pending { considered, next_no_show, uncovered, maximum_broadcast, clock_drift }, then we trigger if our assignment's tranche is less than or equal to maximum_broadcast and the current tick, with clock_drift applied, is at least the tick of our tranche.
    • If we have RequiredTranches::Exact { .. } then we do not trigger, because this value indicates that no new assignments are needed at the moment.
  • If we should trigger our assignment
    • Import the assignment to the ApprovalEntry
    • Broadcast on network with an ApprovalDistributionMessage::DistributeAssignment.
    • Launch approval work for the candidate.
  • Schedule a new wakeup of the candidate.

Schedule Wakeup

  • Requires (approval_entry, candidate_entry) which effectively denotes a (Block Hash, Candidate Hash) pair - the candidate, along with the block it appears in.
  • Also requires RequiredTranches
  • If the approval_entry is approved, this doesn't need to be woken up again.
  • If RequiredTranches::All - no wakeup. We assume other incoming votes will trigger wakeup and potentially re-schedule.
  • If RequiredTranches::Pending { considered, next_no_show, uncovered, maximum_broadcast, clock_drift } - schedule at the lesser of the next no-show tick, or the tick, offset positively by clock_drift of the next non-empty tranche we are aware of after considered, including any tranche containing our own unbroadcast assignment. This can lead to no wakeup in the case that we have already broadcast our assignment and there are no pending no-shows; that is, we have approval votes for every assignment we've received that is not already a no-show. In this case, we will be re-triggered by other validators broadcasting their assignments.
  • If RequiredTranches::Exact { next_no_show, latest_assignment_tick, .. } - set a wakeup for the earlier of the next no-show tick or the latest assignment tick + APPROVAL_DELAY.

Launch Approval Work

  • Requires (SessionIndex, SessionInfo, CandidateReceipt, ValidatorIndex, backing_group, block_hash, candidate_index)
  • Extract the public key of the ValidatorIndex from the SessionInfo for the session.
  • Issue an AvailabilityRecoveryMessage::RecoverAvailableData(candidate, session_index, Some(backing_group), response_sender)
  • Load the historical validation code of the parachain by dispatching a RuntimeApiRequest::ValidationCodeByHash(descriptor.validation_code_hash) against the state of block_hash.
  • Spawn a background task with a clone of background_tx
    • Wait for the available data
    • Issue a CandidateValidationMessage::ValidateFromExhaustive message with APPROVAL_EXECUTION_TIMEOUT as the timeout parameter.
    • Wait for the result of validation
    • Check that the result of validation, if valid, matches the commitments in the receipt.
    • If valid, issue a message on background_tx detailing the request.
    • If any of the data, the candidate, or the commitments are invalid, issue on background_tx a DisputeCoordinatorMessage::IssueLocalStatement with valid = false to initiate a dispute.

Issue Approval Vote

  • Fetch the block entry and candidate entry. Ignore if None - we've probably just lost a race with finality.
  • Construct a SignedApprovalVote with the validator index for the session.
  • Import the checked approval vote. It is "checked" as we've just issued the signature.
  • Construct a IndirectSignedApprovalVote using the information about the vote.
  • Dispatch ApprovalDistributionMessage::DistributeApproval.

Determining Approval of Candidate

Determine Required Tranches

This logic is for inspecting an approval entry that tracks the assignments received, along with information on which assignments have corresponding approval votes. Inspection also involves the current time and expected requirements and is used to help the higher-level code determine the following:

  • Whether to broadcast the local assignment
  • Whether to check that the candidate entry has been completely approved.
  • If the candidate is waiting on approval, when to schedule the next wakeup of the (candidate, block) pair at a point where the state machine could be advanced.

These routines are pure functions which only depend on the environmental state. The expectation is that this determination is re-run every time we attempt to update an approval entry: either when we trigger a wakeup to advance the state machine based on a no-show or our own broadcast, or when we receive further assignments or approvals from the network.

Thus it may be that at some point in time, we consider that tranches 0..X is required to be considered, but as we receive more information, we might require fewer tranches. Or votes that we perceived to be missing and require replacement are filled in and change our view.

Requires (approval_entry, approvals_received, tranche_now, block_tick, no_show_duration, needed_approvals)

#![allow(unused)]
fn main() {
enum RequiredTranches {
  // All validators appear to be required, based on tranches already taken and remaining no-shows.
  All,
  // More tranches required - We're awaiting more assignments.
  Pending {
    /// The highest considered delay tranche when counting assignments.
    considered: DelayTranche,
    /// The tick at which the next no-show, of the assignments counted, would occur.
    next_no_show: Option<Tick>,
    /// The highest tranche to consider when looking to broadcast own assignment.
    /// This should be considered along with the clock drift to avoid broadcasting
    /// assignments that are before the local time.
    maximum_broadcast: DelayTranche,
    /// The clock drift, in ticks, to apply to the local clock when determining whether
    /// to broadcast an assignment or when to schedule a wakeup. The local clock should be treated
    /// as though it is `clock_drift` ticks earlier.
    clock_drift: Tick,
  },
  // An exact number of required tranches and a number of no-shows. This indicates that the amount of `needed_approvals` are assigned and additionally all no-shows are covered.
  Exact {
    /// The tranche to inspect up to.
    needed: DelayTranche,
    /// The amount of missing votes that should be tolerated.
    tolerated_missing: usize,
    /// When the next no-show would be, if any. This is used to schedule the next wakeup in the
    /// event that there are some assignments that don't have corresponding approval votes. If this
    /// is `None`, all assignments have approvals.
    next_no_show: Option<Tick>,
    /// The last tick at which a needed assignment was received.
    last_assignment_tick: Option<Tick>,
  }
}
}

Clock-drift and Tranche-taking

Our vote-counting procedure depends heavily on how we interpret time based on the presence of no-shows - assignments which have no corresponding approval after some time.

We have this is because of how we handle no-shows: we keep track of the depth of no-shows we are covering.

As an example: there may be initial no-shows in tranche 0. It'll take no_show_duration ticks before those are considered no-shows. Then, we don't want to immediately take no_show_duration more tranches. Instead, we want to take one tranche for each uncovered no-show. However, as we take those tranches, there may be further no-shows. Since these depth-1 no-shows should have only been triggered after the depth-0 no-shows were already known to be no-shows, we need to discount the local clock by no_show_duration to see whether these should be considered no-shows or not. There may be malicious parties who broadcast their assignment earlier than they were meant to, who shouldn't be counted as instant no-shows. We continue onwards to cover all depth-1 no-shows which may lead to depth-2 no-shows and so on.

Likewise, when considering how many tranches to take, the no-show depth should be used to apply a depth-discount or clock drift to the tranche_now.

Procedure

  • Start with depth = 0.
  • Set a clock drift of depth * no_show_duration
  • Take tranches up to tranche_now - clock_drift until all needed assignments are met.
  • Keep track of the next_no_show according to the clock drift, as we go.
  • Keep track of the last_assignment_tick as we go.
  • If running out of tranches before then, return Pending { considered, next_no_show, maximum_broadcast, clock_drift }
  • If there are no no-shows, return Exact { needed, tolerated_missing, next_no_show, last_assignment_tick }
  • maximum_broadcast is either DelayTranche::max_value() at tranche 0 or otherwise by the last considered tranche + the number of uncovered no-shows at this point.
  • If there are no-shows, return to the beginning, incrementing depth and attempting to cover the number of no-shows. Each no-show must be covered by a non-empty tranche, which are tranches that have at least one assignment. Each non-empty tranche covers exactly one no-show.
  • If at any point, it seems that all validators are required, do an early return with RequiredTranches::All which indicates that everyone should broadcast.

Check Approval

  • Check whether a candidate is approved under a particular block.
  • Requires (block_entry, candidate_entry, approval_entry, n_tranches)
  • If we have 3 * n_approvals > n_validators, return true. This is because any set with f+1 validators must have at least one honest validator, who has approved the candidate.
  • If n_tranches is RequiredTranches::Pending, return false
  • If n_tranches is RequiredTranches::All, return false.
  • If n_tranches is RequiredTranches::Exact { tranche, tolerated_missing, latest_assignment_tick, .. }, then we return whether all assigned validators up to tranche less tolerated_missing have approved and latest_assignment_tick + APPROVAL_DELAY >= tick_now.
    • e.g. if we had 5 tranches and 1 tolerated missing, we would accept only if all but 1 of assigned validators in tranches 0..=5 have approved. In that example, we also accept all validators in tranches 0..=5 having approved, but that would indicate that the RequiredTranches value was incorrectly constructed, so it is not realistic. tolerated_missing actually represents covered no-shows. If there are more missing approvals than there are tolerated missing, that indicates that there are some assignments which are not yet no-shows, but may become no-shows, and we should wait for the validators to either approve or become no-shows.
    • e.g. If the above passes and the latest_assignment_tick was 5 and the current tick was 6, then we'd return false.

Time

Current Tranche

  • Given the slot number of a block, and the current time, this informs about the current tranche.
  • Convert time.saturating_sub(slot_number.to_time()) to a delay tranches value

Approval Distribution

A subsystem for the distribution of assignments and approvals for approval checks on candidates over the network.

The Approval Voting subsystem is responsible for active participation in a protocol designed to select a sufficient number of validators to check each and every candidate which appears in the relay chain. Statements of participation in this checking process are divided into two kinds:

  • Assignments indicate that validators have been selected to do checking
  • Approvals indicate that validators have checked and found the candidate satisfactory.

The Approval Voting subsystem handles all the issuing and tallying of this protocol, but this subsystem is responsible for the disbursal of statements among the validator-set.

The inclusion pipeline of candidates concludes after availability, and only after inclusion do candidates actually get pushed into the approval checking pipeline. As such, this protocol deals with the candidates made available by particular blocks, as opposed to the candidates which actually appear within those blocks, which are the candidates backed by those blocks. Unless stated otherwise, whenever we reference a candidate partially by block hash, we are referring to the set of candidates made available by those blocks.

We implement this protocol as a gossip protocol, and like other parachain-related gossip protocols our primary concerns are about ensuring fast message propagation while maintaining an upper bound on the number of messages any given node must store at any time.

Approval messages should always follow assignments, so we need to be able to discern two pieces of information based on our View:

  1. Is a particular assignment relevant under a given View?
  2. Is a particular approval relevant to any assignment in a set?

For our own local view, these two queries must not yield false negatives. When applied to our peers' views, it is acceptable for them to yield false negatives. The reason for that is that our peers' views may be beyond ours, and we are not capable of fully evaluating them. Once we have caught up, we can check again for false negatives to continue distributing.

For assignments, what we need to be checking is whether we are aware of the (block, candidate) pair that the assignment references. For approvals, we need to be aware of an assignment by the same validator which references the candidate being approved.

However, awareness on its own of a (block, candidate) pair would imply that even ancient candidates all the way back to the genesis are relevant. We are actually not interested in anything before finality.

We gossip assignments along a grid topology produced by the Gossip Support Subsystem and also to a few random peers. The first time we accept an assignment or approval, regardless of the source, which originates from a validator peer in a shared dimension of the grid, we propagate the message to validator peers in the unshared dimension as well as a few random peers.

But, in case these mechanisms don't work on their own, we need to trade bandwidth for protocol liveness by introducing aggression.

Aggression has 3 levels: Aggression Level 0: The basic behaviors described above. Aggression Level 1: The originator of a message sends to all peers. Other peers follow the rules above. Aggression Level 2: All peers send all messages to all their row and column neighbors. This means that each validator will, on average, receive each message approximately 2*sqrt(n) times.

These aggression levels are chosen based on how long a block has taken to finalize: assignments and approvals related to the unfinalized block will be propagated with more aggression. In particular, it's only the earliest unfinalized blocks that aggression should be applied to, because descendants may be unfinalized only by virtue of being descendants.

Protocol

Input:

  • ApprovalDistributionMessage::NewBlocks
  • ApprovalDistributionMessage::DistributeAssignment
  • ApprovalDistributionMessage::DistributeApproval
  • ApprovalDistributionMessage::NetworkBridgeUpdate
  • OverseerSignal::BlockFinalized

Output:

  • ApprovalVotingMessage::CheckAndImportAssignment
  • ApprovalVotingMessage::CheckAndImportApproval
  • NetworkBridgeMessage::SendValidationMessage::ApprovalDistribution

Functionality

#![allow(unused)]
fn main() {
type BlockScopedCandidate = (Hash, CandidateHash);

enum PendingMessage {
  Assignment(IndirectAssignmentCert, CoreIndex),
  Approval(IndirectSignedApprovalVote),
}

/// The `State` struct is responsible for tracking the overall state of the subsystem.
///
/// It tracks metadata about our view of the unfinalized chain, which assignments and approvals we have seen, and our peers' views.
struct State {
  // These two fields are used in conjunction to construct a view over the unfinalized chain.
  blocks_by_number: BTreeMap<BlockNumber, Vec<Hash>>,
  blocks: HashMap<Hash, BlockEntry>,

  /// Our view updates to our peers can race with `NewBlocks` updates. We store messages received
  /// against the directly mentioned blocks in our view in this map until `NewBlocks` is received.
  ///
  /// As long as the parent is already in the `blocks` map and `NewBlocks` messages aren't delayed
  /// by more than a block length, this strategy will work well for mitigating the race. This is
  /// also a race that occurs typically on local networks.
  pending_known: HashMap<Hash, Vec<(PeerId, PendingMessage>)>>,

  // Peer view data is partially stored here, and partially inline within the `BlockEntry`s
  peer_views: HashMap<PeerId, View>,
}

enum MessageFingerprint {
  Assigment(Hash, u32, ValidatorIndex),
  Approval(Hash, u32, ValidatorIndex),
}

struct Knowledge {
  known_messages: HashSet<MessageFingerprint>,
}

struct PeerKnowledge {
  /// The knowledge we've sent to the peer.
  sent: Knowledge,
  /// The knowledge we've received from the peer.
  received: Knowledge,
}

/// Information about blocks in our current view as well as whether peers know of them.
struct BlockEntry {
  // Peers who we know are aware of this block and thus, the candidates within it. This maps to their knowledge of messages.
  known_by: HashMap<PeerId, PeerKnowledge>,
  // The number of the block.
  number: BlockNumber,
  // The parent hash of the block.
  parent_hash: Hash,
  // Our knowledge of messages.
  knowledge: Knowledge,
  // A votes entry for each candidate.
  candidates: IndexMap<CandidateHash, CandidateEntry>,
}

enum ApprovalState {
  Assigned(AssignmentCert),
  Approved(AssignmentCert, ApprovalSignature),
}

/// Information about candidates in the context of a particular block they are included in. In other words,
/// multiple `CandidateEntry`s may exist for the same candidate, if it is included by multiple blocks - this is likely the case
/// when there are forks.
struct CandidateEntry {
  approvals: HashMap<ValidatorIndex, ApprovalState>,
}
}

Network updates

NetworkBridgeEvent::PeerConnected

Add a blank view to the peer_views state.

NetworkBridgeEvent::PeerDisconnected

Remove the view under the associated PeerId from State::peer_views.

Iterate over every BlockEntry and remove PeerId from it.

NetworkBridgeEvent::OurViewChange

Remove entries in pending_known for all hashes not present in the view. Ensure a vector is present in pending_known for each hash in the view that does not have an entry in blocks.

NetworkBridgeEvent::PeerViewChange

Invoke unify_with_peer(peer, view) to catch them up to messages we have.

We also need to use the view.finalized_number to remove the PeerId from any blocks that it won't be wanting information about anymore. Note that we have to be on guard for peers doing crazy stuff like jumping their finalized_number forward 10 trillion blocks to try and get us stuck in a loop for ages.

One of the safeguards we can implement is to reject view updates from peers where the new finalized_number is less than the previous.

We augment that by defining constrain(x) to output the x bounded by the first and last numbers in state.blocks_by_number.

From there, we can loop backwards from constrain(view.finalized_number) until constrain(last_view.finalized_number) is reached, removing the PeerId from all BlockEntrys referenced at that height. We can break the loop early if we ever exit the bound supplied by the first block in state.blocks_by_number.

NetworkBridgeEvent::PeerMessage

If the block hash referenced by the message exists in pending_known, add it to the vector of pending messages and return.

If the message is of type ApprovalDistributionV1Message::Assignment(assignment_cert, claimed_index), then call import_and_circulate_assignment(MessageSource::Peer(sender), assignment_cert, claimed_index)

If the message is of type ApprovalDistributionV1Message::Approval(approval_vote), then call import_and_circulate_approval(MessageSource::Peer(sender), approval_vote)

Subsystem Updates

ApprovalDistributionMessage::NewBlocks

Create BlockEntry and CandidateEntries for all blocks.

For all entries in pending_known:

  • If there is now an entry under blocks for the block hash, drain all messages and import with import_and_circulate_assignment and import_and_circulate_approval.

For all peers:

  • Compute view_intersection as the intersection of the peer's view blocks with the hashes of the new blocks.
  • Invoke unify_with_peer(peer, view_intersection).

ApprovalDistributionMessage::DistributeAsignment

Call import_and_circulate_assignment with MessageSource::Local.

ApprovalDistributionMessage::DistributeApproval

Call import_and_circulate_approval with MessageSource::Local.

OverseerSignal::BlockFinalized

Prune all lists from blocks_by_number with number less than or equal to finalized_number. Prune all the BlockEntrys referenced by those lists.

Utility

#![allow(unused)]
fn main() {
enum MessageSource {
  Peer(PeerId),
  Local,
}
}

import_and_circulate_assignment(source: MessageSource, assignment: IndirectAssignmentCert, claimed_candidate_index: CandidateIndex)

Imports an assignment cert referenced by block hash and candidate index. As a postcondition, if the cert is valid, it will have distributed the cert to all peers who have the block in their view, with the exclusion of the peer referenced by the MessageSource.

We maintain a few invariants:

  • we only send an assignment to a peer after we add its fingerprint to our knowledge
  • we add a fingerprint of an assignment to our knowledge only if it's valid and hasn't been added before

The algorithm is the following:

  • Load the BlockEntry using assignment.block_hash. If it does not exist, report the source if it is MessageSource::Peer and return.
  • Compute a fingerprint for the assignment using claimed_candidate_index.
  • If the source is MessageSource::Peer(sender):
    • check if peer appears under known_by and whether the fingerprint is in the knowledge of the peer. If the peer does not know the block, report for providing data out-of-view and proceed. If the peer does know the block and the sent knowledge contains the fingerprint, report for providing replicate data and return, otherwise, insert into the received knowledge and return.
    • If the message fingerprint appears under the BlockEntry's Knowledge, give the peer a small positive reputation boost, add the fingerprint to the peer's knowledge only if it knows about the block and return. Note that we must do this after checking for out-of-view and if the peers knows about the block to avoid being spammed. If we did this check earlier, a peer could provide data out-of-view repeatedly and be rewarded for it.
    • Dispatch ApprovalVotingMessage::CheckAndImportAssignment(assignment) and wait for the response.
    • If the result is AssignmentCheckResult::Accepted
      • If the vote was accepted but not duplicate, give the peer a positive reputation boost
      • add the fingerprint to both our and the peer's knowledge in the BlockEntry. Note that we only doing this after making sure we have the right fingerprint.
    • If the result is AssignmentCheckResult::AcceptedDuplicate, add the fingerprint to the peer's knowledge if it knows about the block and return.
    • If the result is AssignmentCheckResult::TooFarInFuture, mildly punish the peer and return.
    • If the result is AssignmentCheckResult::Bad, punish the peer and return.
  • If the source is MessageSource::Local(CandidateIndex)
    • check if the fingerprint appears under the BlockEntry's knowledge. If not, add it.
  • Load the candidate entry for the given candidate index. It should exist unless there is a logic error in the approval voting subsystem.
  • Set the approval state for the validator index to ApprovalState::Assigned unless the approval state is set already. This should not happen as long as the approval voting subsystem instructs us to ignore duplicate assignments.
  • Dispatch a ApprovalDistributionV1Message::Assignment(assignment, candidate_index) to all peers in the BlockEntry's known_by set, excluding the peer in the source, if source has kind MessageSource::Peer. Add the fingerprint of the assignment to the knowledge of each peer.

import_and_circulate_approval(source: MessageSource, approval: IndirectSignedApprovalVote)

Imports an approval signature referenced by block hash and candidate index:

  • Load the BlockEntry using approval.block_hash and the candidate entry using approval.candidate_entry. If either does not exist, report the source if it is MessageSource::Peer and return.
  • Compute a fingerprint for the approval.
  • Compute a fingerprint for the corresponding assignment. If the BlockEntry's knowledge does not contain that fingerprint, then report the source if it is MessageSource::Peer and return. All references to a fingerprint after this refer to the approval's, not the assignment's.
  • If the source is MessageSource::Peer(sender):
    • check if peer appears under known_by and whether the fingerprint is in the knowledge of the peer. If the peer does not know the block, report for providing data out-of-view and proceed. If the peer does know the block and the sent knowledge contains the fingerprint, report for providing replicate data and return, otherwise, insert into the received knowledge and return.
    • If the message fingerprint appears under the BlockEntry's Knowledge, give the peer a small positive reputation boost, add the fingerprint to the peer's knowledge only if it knows about the block and return. Note that we must do this after checking for out-of-view to avoid being spammed. If we did this check earlier, a peer could provide data out-of-view repeatedly and be rewarded for it.
    • Dispatch ApprovalVotingMessage::CheckAndImportApproval(approval) and wait for the response.
    • If the result is VoteCheckResult::Accepted(()):
      • Give the peer a positive reputation boost and add the fingerprint to both our and the peer's knowledge.
    • If the result is VoteCheckResult::Bad:
      • Report the peer and return.
  • Load the candidate entry for the given candidate index. It should exist unless there is a logic error in the approval voting subsystem.
  • Set the approval state for the validator index to ApprovalState::Approved. It should already be in the Assigned state as our BlockEntry knowledge contains a fingerprint for the assignment.
  • Dispatch a ApprovalDistributionV1Message::Approval(approval) to all peers in the BlockEntry's known_by set, excluding the peer in the source, if source has kind MessageSource::Peer. Add the fingerprint of the assignment to the knowledge of each peer. Note that this obeys the politeness conditions:
    • We guarantee elsewhere that all peers within known_by are aware of all assignments relative to the block.
    • We've checked that this specific approval has a corresponding assignment within the BlockEntry.
    • Thus, all peers are aware of the assignment or have a message to them in-flight which will make them so.

unify_with_peer(peer: PeerId, view):

  1. Initialize a set missing_knowledge = {}

For each block in the view: 2. Load the BlockEntry for the block. If the block is unknown, or the number is less than or equal to the view's finalized number go to step 6. 3. Inspect the known_by set of the BlockEntry. If the peer already knows all assignments/approvals, go to step 6. 4. Add the peer to known_by and add the hash and missing knowledge of the block to missing_knowledge. 5. Return to step 2 with the ancestor of the block.

  1. For each block in missing_knowledge, send all assignments and approvals for all candidates in those blocks to the peer.

Disputes Subsystems

If approval voting finds an invalid candidate, a dispute is raised. The disputes subsystems are concerned with the following:

  1. Disputes can be raised
  2. Disputes (votes) get propagated to all other validators
  3. Votes get recorded as necessary
  4. Nodes will participate in disputes in a sensible fashion
  5. Finality is stopped while a candidate is being disputed on chain
  6. Chains can be reverted in case a dispute concludes invalid
  7. Votes are provided to the provisioner for importing on chain, in order for slashing to work.

The dispute-coordinator subsystem interfaces with the provisioner and chain selection to make the bulk of this possible. dispute-distribution is concerned with getting votes out to other validators and receiving them in a spam resilient way.

Dispute Coordinator

The coordinator is the central subsystem of the node-side components which participate in disputes. It wraps a database, which is used to track statements observed by all validators over some window of sessions. Votes older than this session window are pruned.

In particular the dispute-coordinator is responsible for:

  • Ensuring that the node is able to raise a dispute in case an invalid candidate is found during approval checking.
  • Ensuring that backing and approval votes will be recorded on chain. With these votes on chain we can be certain that appropriate targets for slashing will be available for concluded disputes. Also, scraping these votes during a dispute is necessary for critical spam prevention measures.
  • Ensuring backing votes will never get overridden by explicit votes.
  • Coordinating actual participation in a dispute, ensuring that the node participates in any justified dispute in a way that ensures resolution of disputes on the network even in the case of many disputes raised (flood/DoS scenario).
  • Ensuring disputes resolve, even for candidates on abandoned forks as much as reasonably possible, to rule out "free tries" and thus guarantee our gambler's ruin property.
  • Providing an API for chain selection, so we can prevent finalization of any chain which has included candidates for which a dispute is either ongoing or concluded invalid and avoid building on chains with an included invalid candidate.
  • Providing an API for retrieving (resolved) disputes, including all votes, both implicit (approval, backing) and explicit dispute votes. So validators can get rewarded/slashed accordingly.

Ensuring That Disputes Can Be Raised

If a candidate turns out invalid in approval checking, the approval-voting subsystem will try to issue a dispute. For this, it will send a message DisputeCoordinatorMessage::IssueLocalStatement to the dispute coordinator, indicating to cast an explicit invalid vote. It is the responsibility of the dispute coordinator on reception of such a message to create and sign that explicit invalid vote and trigger a dispute if none for that candidate is already ongoing.

In order to raise a dispute, a node has to be able to provide two opposing votes. Given that the reason of the backing phase is to have validators with skin in the game, the opposing valid vote will very likely be a backing vote. It could also be some already cast approval vote, but the significant point here is: As long as we have backing votes available, any node will be able to raise a dispute.

Therefore a vital responsibility of the dispute coordinator is to make sure backing votes are available for all candidates that might still get disputed. To accomplish this task in an efficient way the dispute-coordinator relies on chain scraping. Whenever a candidate gets backed on chain, we record in chain storage the backing votes imported in that block. This way, given the chain state for a given relay chain block, we can retrieve via a provided runtime API the backing votes imported by that block. The dispute coordinator makes sure to query those votes for any non finalized blocks: In case of missed blocks, it will do chain traversal as necessary.

Relying on chain scraping is very efficient for two reasons:

  1. Votes are already batched. We import all available backing votes for a candidate all at once. If instead we imported votes from candidate-backing as they came along, we would import each vote individually which is inefficient in the current dispute coordinator implementation (quadratic complexity).
  2. We also import less votes in total, as we avoid importing statements for candidates that never got successfully backed on any chain.

It also is secure, because disputes are only ever raised in the approval voting phase. A node only starts the approval process after it has seen a candidate included on some chain, for that to happen it must have been backed previously. Therefore backing votes are available at that point in time. Signals are processed first, so even if a block is skipped and we only start importing backing votes on the including block, we will have seen the backing votes by the time we process messages from approval voting.

In summary, for making it possible for a dispute to be raised, recording of backing votes from chain is sufficient and efficient. In particular there is no need to preemptively import approval votes, which has shown to be a very inefficient process. (Quadratic complexity adds up, with 35 votes in total per candidate)

Approval votes are very relevant nonetheless as we are going to see in the next section.

Ensuring approval votes will be recorded

Ensuring Recording

Only votes recorded by the dispute coordinator will be considered for slashing.

While there is no need to record approval votes in the dispute coordinator preemptively, we make some effort to have any in approval-voting received approval votes recorded when a dispute actually happens:

This is not required for concluding the dispute, as nodes send their own vote anyway (either explicit valid or their existing approval-vote). What nodes can do though, is participating in approval-voting, casting a vote, but later when a dispute is raised reconsider their vote and send an explicit invalid vote. If they managed to only have that one recorded, then they could avoid a slash.

This is not a problem for our basic security assumptions: The backers are the ones to be supposed to have skin in the game, so we are not too woried about colluding approval voters getting away slash free as the gambler's ruin property is maintained anyway. There is however a separate problem, from colluding approval-voters, that is "lazy" approval voters. If it were easy and reliable for approval-voters to reconsider their vote, in case of an actual dispute, then they don't have a direct incentive (apart from playing a part in securing the network) to properly run the validation function at all - they could just always vote "valid" totally risk free. (While they would alwasy risk a slash by voting invalid.)

So we do want to fetch approval votes from approval-voting. Importing votes is most efficient when batched. At the same time approval voting and disputes are running concurrently so approval votes are expected to trickle in still, when a dispute is already ongoing.

Hence, we have the following requirements for importing approval votes:

  1. Only import them when there is a dispute, because otherwise we are wasting lots of resources always for the exceptional case of a dispute.
  2. Import votes batched when possible, to avoid quadratic import complexity.
  3. Take into account that approval voting is still ongoing, while a dispute is already running.

With a design where approval voting sends votes to the dispute-coordinator by itself, we would need to make approval voting aware of ongoing disputes and once it is aware it could start sending all already existing votes batched and trickling in votes as they come. The problem with this is, that it adds some unnecessary complexity to approval-voting and also we might still import most of the votes unbatched one-by-one, depending on what point in time the dispute was raised.

Instead of the dispute coordinator informing approval-voting of an ongoing dispute for it to begin forwarding votes to the dispute coordinator, it makes more sense for the dispute-coordinator to just ask approval-voting for votes of candidates in dispute. This way, the dispute coordinator can also pick the best time for maximizing the number of votes in the batch.

Now the question remains, when should the dispute coordinator ask approval-voting for votes?

In fact for slashing it is only relevant to have them once the dispute concluded, so we can query approval voting the moment the dispute concludes! Two concerns that come to mind, are easily addressed:

  1. Timing: We would like to rely as little as possible on implementation details of approval voting. In particular, if the dispute is ongoing for a long time, do we have any guarantees that approval votes are kept around long enough by approval voting? Will approval votes still be present by the time the dispute concludes in all cases? The answer is nuanced, but in general we cannot rely on it. The problem is first, that finalization and approval-voting is an off-chain process so there is no global consensus: As soon as at least f+1 honest (f=n/3, where n is the number of validators/nodes) nodes have seen the dispute conclude, finalization will take place and approval votes will be cleared. This would still be fine, if we had some guarantees that those honest nodes will be able to include those votes in a block. This guarantee does not exist unfortunately, we will discuss the problem and solutions in more detail [below][#Ensuring Chain Import].

    The second problem is that approval-voting will abandon votes as soon as a chain can no longer be finalized (some other/better fork already has been). This second problem can somehow be mitigated by also importing votes as soon as a dispute is detected, but not fully resolved. It is still inherently racy. The good thing is, this should be good enough: We are worried about lazy approval checkers, the system does not need to be perfect. It should be enough if there is some risk of getting caught.

  2. We are not worried about the dispute not concluding, as nodes will always send their own vote, regardless of it being an explict or an already existing approval-vote.

Conclusion: As long as we make sure, if our own approval vote gets imported (which would prevent dispute participation) to also distribute it via dispute-distribution, disputes can conclude. To mitigate raciness with approval-voting deleting votes we will import approval votes twice during a dispute: Once when it is raised, to make as sure as possible to see approval votes also for abandoned forks and second when the dispute concludes, to maximize the amount of potentially malicious approval votes to be recorded. The raciness obviously is not fully resolved by this, but this is fine as argued above.

Ensuring vote import on chain is covered in the next section.

What we don't care about is that honest approval-voters will likely validate twice, once in approval voting and once via dispute-participation. Avoiding that does not really seem worthwhile though, as disputes are for one exceptional, so a little wasted effort won't affect everyday performance - second, even with eager importing of approval votes, those doubled work is still present as disputes and approvals are racing. Every time participation is faster than approval, a node would do double work.

Ensuring Chain Import

While in the previous section we discussed means for nodes to ensure relevant votes are recorded so lazy approval checkers get slashed properly, it is crucial to also discuss the actual chain import. Only if we guarantee that recorded votes will get imported on chain (on all potential chains really) we will succeed in executing slashes. Particularly we need to make sure backing votes end up on chain consistently.

Dispute distribution will make sure all explicit dispute votes get distributed among nodes which includes current block producers (current authority set) which is an important property: If the dispute carries on across an era change, we need to ensure that the new validator set will learn about any disputes and their votes, so they can put that information on chain. Dispute-distribution luckily has this property and always sends votes to the current authority set. The issue is, for dispute-distribution, nodes send only their own explicit (or in some cases their approval vote) in addition to some opposing vote. This guarantees that at least some backing or approval vote will be present at the block producer, but we don't have a 100% guarantee to have votes for all backers, even less for approval checkers.

Reason for backing votes: While backing votes will be present on at least some chain, that does not mean that any such chain is still considered for block production in the current set - they might only exist on an already abandoned fork. This means a block producer that just joined the set, might not have seen any of them.

For approvals it is even more tricky and less necessary: Approval voting together with finalization is a completely off-chain process therefore those protocols don't care about block production at all. Approval votes only have a guarantee of being propagated between the nodes that are responsible for finalizing the concerned blocks. This implies that on an era change the current authority set, will not necessarily get informed about any approval votes for the previous era. Hence even if all validators of the previous era successfully recorded all approval votes in the dispute coordinator, they won't get a chance to put them on chain, hence they won't be considered for slashing.

It is important to note, that the essential properties of the system still hold: Dispute-distribution will distribute at least one "valid" vote to the current authority set, hence at least one node will get slashed in case of outcome "invalid". Also in reality the validator set is rarely exchanged 100%, therefore in practice some validators in the current authority set will overlap with the ones in the previous set and will be able to record votes on chain.

Still, for maximum accountability we need to make sure a previous authority set can communicate votes to the next one, regardless of any chain: This is yet to be implemented see section "Resiliency" in dispute-distribution and this ticket.

Coordinating Actual Dispute Participation

Once the dispute coordinator learns about a dispute, it is its responsibility to make sure the local node participates in that dispute.

The dispute coordinator learns about a dispute by importing votes from either chain scraping or from dispute-distribution. If it finds opposing votes (always the case when coming from dispute-distribution), it records the presence of a dispute. Then, in case it does not find any local vote for that dispute already, it needs to trigger participation in the dispute (see previous section for considerations when the found local vote is an approval vote).

Participation means, recovering availability and re-evaluating the POV. The result of that validation (either valid or invalid) will be the node's vote on that dispute: Either explicit "invalid" or "valid". The dispute coordinator will inform dispute-distribution about our vote and dispute-distribution will make sure that our vote gets distributed to all other validators.

Nothing ever is that easy though. We can not blindly import anything that comes along and trigger participation no matter what.

Spam Considerations

In Polkadot's security model, it is important that attempts to attack the system result in a slash of the offenders. Therefore we need to make sure that this slash is actually happening. Attackers could try to prevent the slashing from taking place, by overwhelming validators with disputes in such a way that no single dispute ever concludes, because nodes are busy processing newly incoming ones. Other attacks are imaginable as well, like raising disputes for candidates that don't exist, just filling up everyone's disk slowly or worse making nodes try to participate, which will result in lots of network requests for recovering availability.

The last point brings up a significant consideration in general: Disputes are about escalation: Every node will suddenly want to check, instead of only a few. A single message will trigger the whole network to start significant amount of work and will cause lots of network traffic and messages. Hence the dispute system is very susceptible to being a brutal amplifier for DoS attacks, resulting in DoS attacks to become very easy and cheap, if we are not careful.

One counter measure we are taking is making raising of disputes a costly thing: If you raise a dispute, because you claim a candidate is invalid, although it is in fact valid - you will get slashed, hence you pay for consuming those resources. The issue is: This only works if the dispute concerns a candidate that actually exists!

If a node raises a dispute for a candidate that never got included (became available) on any chain, then the dispute can never conclude, hence nobody gets slashed. It makes sense to point out that this is less bad than it might sound at first, as trying to participate in a dispute for a non existing candidate is "relatively" cheap. Each node will send out a few hundred tiny request messages for availability chunks, which all will end up in a tiny response "NoSuchChunk" and then no participation will actually happen as there is nothing to participate. Malicious nodes could provide chunks, which would make things more costly, but at the full expense of the attackers bandwidth - no amplification here. I am bringing that up for completeness only: Triggering a thousand nodes to send out a thousand tiny network messages by just sending out a single garbage message, is still a significant amplification and is nothing to ignore - this could absolutely be used to cause harm!

Participation

As explained, just blindly participating in any "dispute" that comes along is not a good idea. First we would like to make sure the dispute is actually genuine, to prevent cheap DoS attacks. Secondly, in case of genuine disputes, we would like to conclude one after the other, in contrast to processing all at the same time, slowing down progress on all of them, bringing individual processing to a complete halt in the worst case (nodes get overwhelmed at some stage in the pipeline).

To ensure to only spend significant work on genuine disputes, we only trigger participation at all on any vote import if any of the following holds true:

  • We saw the disputed candidate included in some not yet finalized block on at least one fork of the chain.
  • We have seen the disputed candidate backed in some not yet finalized block on at least one fork of the chain. This ensures the candidate is at least not completely made up and there has been some effort already flown into that candidate. Generally speaking a dispute shouldn't be raised for a candidate which is backed but is not yet included. Disputes are raised during approval checking. We participate on such disputes as a precaution - maybe we haven't seen the CandidateIncluded event yet?
  • The dispute is already confirmed: Meaning that 1/3+1 nodes already participated, as this suggests in our threat model that there was at least one honest node that already voted, so the dispute must be genuine.

Note: A node might be out of sync with the chain and we might only learn about a block, including a candidate, after we learned about the dispute. This means, we have to re-evaluate participation decisions on block import!

With this, nodes won't waste significant resources on completely made up candidates. The next step is to process dispute participation in a (globally) ordered fashion. Meaning a majority of validators should arrive at at least roughly at the same ordering of participation, for disputes to get resolved one after another. This order is only relevant if there are lots of disputes, so we obviously only need to worry about order if participations start queuing up.

We treat participation for candidates that we have seen included with priority and put them on a priority queue which sorts participation based on the block number of the relay parent of the candidate and for candidates with the same relay parent height further by the CandidateHash. This ordering is globally unique and also prioritizes older candidates.

The latter property makes sense, because if an older candidate turns out invalid, we can roll back the full chain at once. If we resolved earlier disputes first and they turned out invalid as well, we might need to roll back a couple of times instead of just once to the oldest offender. This is obviously a good idea, in particular it makes it impossible for an attacker to prevent rolling back a very old candidate, by keeping raising disputes for newer candidates.

For candidates we have not seen included, but we know are backed (thanks to chain scraping) or we have seen a dispute with 1/3+1 participation (confirmed dispute) on them - we put participation on a best-effort queue. It has got the same ordering as the priority one - by block heights of the relay parent, older blocks are with priority. There is a possibility not to be able to obtain the block number of the parent when we are inserting the dispute in the queue. To account for races, we will promote any existing participation request to the priority queue once we learn about an including block. NOTE: this is still work in progress and is tracked by this issue.

Abandoned Forks

Finalization: As mentioned we care about included and backed candidates on any non-finalized chain, given that any disputed chain will not get finalized, we don't need to care about finalized blocks, but what about forks that fall behind the finalized chain in terms of block number? For those we would still like to be able to participate in any raised disputes, otherwise attackers might be able to avoid a slash if they manage to create a better fork after they learned about the approval checkers. Therefore we do care about those forks even after they have fallen behind the finalized chain.

For simplicity we also care about the actual finalized chain (not just forks) up to a certain depth. We do have to limit the depth, because otherwise we open a DoS vector again. The depth (into the finalized chain) should be oriented on the approval-voting execution timeout, in particular it should be significantly larger. Otherwise by the time the execution is allowed to finish, we already dropped information about those candidates and the dispute could not conclude.

Import

Spam Considerations

In the last section we looked at how to treat queuing participations to handle heavy dispute load well. This already ensures, that honest nodes won't amplify cheap DoS attacks. There is one minor issue remaining: Even if we delay participation until we have some confirmation of the authenticity of the dispute, we should also not blindly import all votes arriving into the database as this might be used to just slowly fill up disk space, until the node is no longer functional. This leads to our last protection mechanism at the dispute coordinator level (dispute-distribution also has its own), which is spam slots. For each import containing an invalid vote, where we don't know whether it might be spam or not we increment a counter for each signing participant of explicit invalid votes.

What votes do we treat as a potential spam? A vote will increase a spam slot if and only if all of the following conditions are satisfied:

  • the candidate under dispute was not seen included nor backed on any chain
  • the dispute is not confirmed
  • we haven't cast a vote for the dispute

Whenever any vote on a dispute is imported these conditions are checked. If the dispute is found not to be potential spam, then spam slots for the disputed candidate hash are cleared. This decrements the spam count for every validator which had voted invalid.

To keep spam slots from filling up unnecessarily we want to clear spam slots whenever a candidate is seen to be backed or included. Fortunately this behavior is acheived by clearing slots on vote import as described above. Because on chain backing votes are processed when a block backing the disputed candidate is discovered, spam slots are cleared for every backed candidate. Included candidates have also been seen as backed on the same fork, so decrementing spam slots is handled in that case as well.

The reason this works is because we only need to worry about actual dispute votes. Import of backing votes are already rate limited and concern only real candidates. For approval votes a similar argument holds (if they come from approval-voting), but we also don't import them until a dispute already concluded. For actual dispute votes we need two opposing votes, so there must be an explicit invalid vote in the import. Only a third of the validators can be malicious, so spam disk usage is limited to 2*vote_size*n/3*NUM_SPAM_SLOTS, with n being the number of validators.

Backing Votes

Backing votes are in some way special. For starters they are the only valid votes that are guaranteed to exist for any valid dispute to be raised. Second they are the only votes that commit to a shorter execution timeout BACKING_EXECUTION_TIMEOUT, compared to a more lenient timeout used in approval voting. To account properly for execution time variance across machines, slashing might treat backing votes differently (more aggressively) than other voting valid votes. Hence in import we shall never override a backing vote with another valid vote. They can not be assumed to be interchangeable.

Attacks & Considerations

The following attacks on the priority queue and best-effort queues are considered in above design.

Priority Queue

On the priority queue, we will only queue participations for candidates we have seen included on any chain. Any attack attempt would start with a candidate included on some chain, but an attacker could try to only reveal the including relay chain blocks to just some honest validators and stop as soon as it learns that some honest validator would have a relevant approval assignment.

Without revealing the including block to any honest validator, we don't really have an attack yet. Once the block is revealed though, the above is actually very hard. Each honest validator will re-distribute the block it just learned about. This means an attacker would need to pull of a targeted DoS attack, which allows the validator to send its assignment, but prevents it from forwarding and sharing the relay chain block.

This sounds already hard enough, provided that we also start participation if we learned about an including block after the dispute has been raised already (we need to update participation queues on new leaves), but to be even safer we choose to have an additional best-effort queue.

Best-Effort Queue

While attacking the priority queue is already pretty hard, attacking the best-effort queue is even harder. For a candidate to be a threat, it has to be included on some chain. For it to be included, it has to have been backed before and at least n/3 honest nodes must have seen that block, so availability (inclusion) can be reached. Making a full third of the nodes not further propagate a block, while at the same time allowing them to fetch chunks, sign and distribute bitfields seems almost infeasible and even if accomplished, those nodes would be enough to confirm a dispute and we have not even touched the above fact that in addition, for an attack, the following including block must be shared with honest validators as well.

It is worth mentioning that a successful attack on the priority queue as outlined above is already outside of our threat model, as it assumes n/3 malicious nodes + additionally malfunctioning/DoSed nodes. Even more so for attacks on the best-effort queue, as our threat model only allows for n/3 malicious or malfunctioning nodes in total. It would therefore be a valid decision to ditch the best-effort queue, if it proves to become a burden or creates other issues.

One issue we should not be worried about though is spam. For abusing best-effort for spam, the following scenario would be necessary:

An attacker controls a backing group: The attacker can then have candidates backed and choose to not provide chunks. This should come at a cost to miss out on rewards for backing, so is not free. At the same time it is rate limited, as a backing group can only back so many candidates legitimately. (~ 1 per slot):

  1. They have to wait until a malicious actor becomes block producer (for causing additional forks via equivocation for example).
  2. Forks are possible, but if caused by equivocation also not free.
  3. For each fork the attacker has to wait until the candidate times out, for backing another one.

Assuming there can only be a handful of forks, 2) together with 3) the candidate timeout restriction, frequency should indeed be in the ballpark of once per slot. Scaling linearly in the number of controlled backing groups, so two groups would mean 2 backings per slot, ...

So by this reasoning an attacker could only do very limited harm and at the same time will have to pay some price for it (it will miss out on rewards). Overall the work done by the network might even be in the same ballpark as if actors just behaved honestly:

  1. Validators would have fetched chunks
  2. Approval checkers would have done approval checks

While because of the attack (backing, not providing chunks and afterwards disputing the candidate), the work for 1000 validators would be:

All validators sending out ~ 1000 tiny requests over already established connections, with also tiny (byte) responses.

This means around a million requests, while in the honest case it would be ~ 10000 (30 approval checkers x330) - where each request triggers a response in the range of kilobytes. Hence network load alone will likely be higher in the honest case than in the DoS attempt case, which would mean the DoS attempt actually reduces load, while also costing rewards.

In the worst case this can happen multiple times, as we would retry that on every vote import. The effect would still be in the same ballpark as honest behavior though and can also be mitigated by chilling repeated availability recovery requests for example.

Out of Scope

No Disputes for Non Included Candidates

We only ever care about disputes for candidates that have been included on at least some chain (became available). This is because the availability system was designed for precisely that: Only with inclusion (availability) we have guarantees about the candidate to actually be available. Because only then we have guarantees that malicious backers can be reliably checked and slashed. Also, by design non included candidates do not pose any threat to the system.

One could think of an (additional) dispute system to make it possible to dispute any candidate that has been proposed by a validator, no matter whether it got successfully included or even backed. Unfortunately, it would be very brittle (no availability) and also spam protection would be way harder than for the disputes handled by the dispute-coordinator. In fact, all the spam handling strategies described above would simply be unavailable.

It is worth thinking about who could actually raise such disputes anyway: Approval checkers certainly not, as they will only ever check once availability succeeded. The only other nodes that meaningfully could/would are honest backing nodes or collators. For collators spam considerations would be even worse as there can be an unlimited number of them and we can not charge them for spam, so trying to handle disputes raised by collators would be even more complex. For honest backers: It actually makes more sense for them to wait until availability is reached as well, as only then they have guarantees that other nodes will be able to check. If they disputed before, all nodes would need to recover the data from them, so they would be an easy DoS target.

In summary: The availability system was designed for raising disputes in a meaningful and secure way after availability was reached. Trying to raise disputes before does not meaningfully contribute to the systems security/might even weaken it as attackers are warned before availability is reached, while at the same time adding signficant amount of complexity. We therefore punt on such disputes and concentrate on disputes the system was designed to handle.

No Disputes for Already Finalized Blocks

Note that by above rules in the Participation section, we will not participate in disputes concerning a candidate in an already finalized block. This is because, disputing an already finalized block is simply too late and therefore of little value. Once finalized, bridges have already processed the block for example, so we have to assume the damage is already done. Governance has to step in and fix what can be fixed.

Making disputes for already finalized blocks possible would only provide two features:

  1. We can at least still slash attackers.
  2. We can freeze the chain to some governance only mode, in an attempt to minimize potential harm done.

Both seem kind of worthwhile, although as argued above, it is likely that there is not too much that can be done in 2 and we would likely only ending up DoSing the whole system without much we can do. 1 can also be achieved via governance mechanisms.

In any case, our focus should be making as sure as reasonably possible that any potentially invalid block does not get finalized in the first place. Not allowing disputing already finalized blocks actually helps a great deal with this goal as it massively reduces the amount of candidates that can be disputed.

This makes attempts to overwhelm the system with disputes significantly harder and counter measures way easier. We can limit inclusion for example (as suggested here in case of high dispute load. Another measure we have at our disposal is that on finality lag block production will slow down, implicitly reducing the rate of new candidates that can be disputed. Hence, the cutting-off of the unlimited candidate supply of already finalized blocks, guarantees the necessary DoS protection and ensures we can have measures in place to keep up with processing of disputes.

If we allowed participation for disputes for already finalized candidates, the above spam protection mechanisms would be insufficient/relying 100% on full and quick disabling of spamming validators.

Database Schema

We use an underlying Key-Value database where we assume we have the following operations available:

  • write(key, value)
  • read(key) -> Option<value>
  • iter_with_prefix(prefix) -> Iterator<(key, value)> - gives all keys and values in lexicographical order where the key starts with prefix.

We use this database to encode the following schema:

#![allow(unused)]
fn main() {
("candidate-votes", SessionIndex, CandidateHash) -> Option<CandidateVotes>
"recent-disputes" -> RecentDisputes
"earliest-session" -> Option<SessionIndex>
}

The meta information that we track per-candidate is defined as the CandidateVotes struct. This draws on the dispute statement types

#![allow(unused)]
fn main() {
/// Tracked votes on candidates, for the purposes of dispute resolution.
pub struct CandidateVotes {
  /// The receipt of the candidate itself.
  pub candidate_receipt: CandidateReceipt,
  /// Votes of validity, sorted by validator index.
  pub valid: Vec<(ValidDisputeStatementKind, ValidatorIndex, ValidatorSignature)>,
  /// Votes of invalidity, sorted by validator index.
  pub invalid: Vec<(InvalidDisputeStatementKind, ValidatorIndex, ValidatorSignature)>,
}

/// The mapping for recent disputes; any which have not yet been pruned for being ancient.
pub type RecentDisputes = std::collections::BTreeMap<(SessionIndex, CandidateHash), DisputeStatus>;

/// The status of dispute. This is a state machine which can be altered by the
/// helper methods.
pub enum DisputeStatus {
  /// The dispute is active and unconcluded.
  Active,
  /// The dispute has been concluded in favor of the candidate
  /// since the given timestamp.
  ConcludedFor(Timestamp),
  /// The dispute has been concluded against the candidate
  /// since the given timestamp.
  ///
  /// This takes precedence over `ConcludedFor` in the case that
  /// both are true, which is impossible unless a large amount of
  /// validators are participating on both sides.
  ConcludedAgainst(Timestamp),
  /// Dispute has been confirmed (more than `byzantine_threshold` have already participated/ or
  /// we have seen the candidate included already/participated successfully ourselves).
  Confirmed,
}
}

Protocol

Input: DisputeCoordinatorMessage

Output:

Functionality

This assumes a constant DISPUTE_WINDOW: SessionWindowSize. This should correspond to at least 1 day.

Ephemeral in-memory state:

#![allow(unused)]
fn main() {
struct State {
  keystore: Arc<LocalKeystore>,
  rolling_session_window: RollingSessionWindow,
  highest_session: SessionIndex,
  spam_slots: SpamSlots,
  participation: Participation,
  ordering_provider: OrderingProvider,
  participation_receiver: WorkerMessageReceiver,
  metrics: Metrics,
  // This tracks only rolling session window failures.
  // It can be a `Vec` if the need to track more arises.
  error: Option<SessionsUnavailable>,
  /// Latest relay blocks that have been successfully scraped.
  last_scraped_blocks: LruCache<Hash, ()>,
}
}

On startup

When the subsystem is initialised it waits for a new leaf (message OverseerSignal::ActiveLeaves). The leaf is used to initialise a RollingSessionWindow instance (contains leaf hash and DISPUTE_WINDOW which is a constant).

Next the active disputes are loaded from the DB and initialize spam slots accordingly, then for each loaded dispute, we either send a DisputeDistribution::SendDispute if there is a local vote from us available or if there is none and participation is in order, we push the dispute to participation.

The main loop

Just after the subsystem initialisation the main loop (fn run_until_error()) runs until OverseerSignal::Conclude signal is received. Before executing the actual main loop the leaf and the participations, obtained during startup are enqueued for processing. If there is capacity (the number of running participations is less than MAX_PARALLEL_PARTICIPATIONS) participation jobs are started (func participate). Finally the component waits for messages from Overseer. The behaviour on each message is described in the following subsections.

On OverseerSignal::ActiveLeaves

Initiates processing via the Participation module and updates the internal state of the subsystem. More concretely:

  • Passes the ActiveLeavesUpdate message to the ordering provider.
  • Updates the session info cache.
  • Updates self.highest_session.
  • Prunes old spam slots in case the session window has advanced.
  • Scrapes on chain votes.

On MuxedMessage::Participation

This message is sent from Participatuion module and indicates a processed dispute participation. It's the result of the processing job initiated with OverseerSignal::ActiveLeaves. The subsystem issues a DisputeMessage with the result.

On OverseerSignal::Conclude

Exit gracefully.

On OverseerSignal::BlockFinalized

Performs cleanup of the finalized candidate.

On DisputeCoordinatorMessage::ImportStatements

Import statements by validators are processed in fn handle_import_statements(). The function has got three main responsibilities:

  • Initiate participation in disputes and sending out of any existing own approval vote in case of a raised dispute.
  • Persist all fresh votes in the database. Fresh votes in this context means votes that are not already processed by the node.
  • Spam protection on all invalid (DisputeStatement::Invalid) votes. Please check the SpamSlots section for details on how spam protection works.

On DisputeCoordinatorMessage::RecentDisputes

Returns all recent disputes saved in the DB.

On DisputeCoordinatorMessage::ActiveDisputes

Returns all recent disputes concluded within the last ACTIVE_DURATION_SECS .

On DisputeCoordinatorMessage::QueryCandidateVotes

Loads candidate-votes for every (SessionIndex, CandidateHash) in the input query and returns data within each CandidateVote. If a particular candidate-vote is missing, that particular request is omitted from the response.

On DisputeCoordinatorMessage::IssueLocalStatement

Executes fn issue_local_statement() which performs the following operations:

  • Deconstruct into parts { session_index, candidate_hash, candidate_receipt, is_valid }.
  • Construct a DisputeStatement based on Valid or Invalid, depending on the parameterization of this routine.
  • Sign the statement with each key in the SessionInfo's list of parachain validation keys which is present in the keystore, except those whose indices appear in voted_indices. This will typically just be one key, but this does provide some future-proofing for situations where the same node may run on behalf multiple validators. At the time of writing, this is not a use-case we support as other subsystems do not invariably provide this guarantee.
  • Write statement to DB.
  • Send a DisputeDistributionMessage::SendDispute message to get the vote distributed to other validators.

On DisputeCoordinatorMessage::DetermineUndisputedChain

Executes fn determine_undisputed_chain() which performs the following:

  • Load "recent-disputes".
  • Deconstruct into parts { base_number, block_descriptions, rx }
  • Starting from the beginning of block_descriptions:
    1. Check the RecentDisputes for a dispute of each candidate in the block description.
    2. If there is a dispute which is active or concluded negative, exit the loop.
  • For the highest index i reached in the block_descriptions, send (base_number + i + 1, block_hash) on the channel, unless i is 0, in which case None should be sent. The block_hash is determined by inspecting block_descriptions[i].

Dispute Distribution

Dispute distribution is responsible for ensuring all concerned validators will be aware of a dispute and have the relevant votes.

Design Goals

This design should result in a protocol that is:

  • resilient to nodes being temporarily unavailable
  • make sure nodes are aware of a dispute quickly
  • relatively efficient, should not cause too much stress on the network
  • be resilient when it comes to spam
  • be simple and boring: We want disputes to work when they happen

Protocol

Distributing disputes needs to be a reliable protocol. We would like to make as sure as possible that our vote got properly delivered to all concerned validators. For this to work, this subsystem won't be gossip based, but instead will use a request/response protocol for application level confirmations. The request will be the payload (the actual votes/statements), the response will be the confirmation. See [below][#wire-format].

Input

DisputeDistributionMessage

Output

  • [DisputeCoordinatorMessage::ActiveDisputes][DisputeCoordinatorMessage]
  • [DisputeCoordinatorMessage::ImportStatements][DisputeCoordinatorMessage]
  • [DisputeCoordinatorMessage::QueryCandidateVotes][DisputeCoordinatorMessage]
  • RuntimeApiMessage

Wire format

Disputes

Protocol: "/<genesis_hash>/<fork_id>/send_dispute/1"

Request:

#![allow(unused)]
fn main() {
struct DisputeRequest {
  /// The candidate being disputed.
  pub candidate_receipt: CandidateReceipt,

  /// The session the candidate appears in.
  pub session_index: SessionIndex,

  /// The invalid vote data that makes up this dispute.
  pub invalid_vote: InvalidDisputeVote,

  /// The valid vote that makes this dispute request valid.
  pub valid_vote: ValidDisputeVote,
}

/// Any invalid vote (currently only explicit).
pub struct InvalidDisputeVote {
  /// The voting validator index.
  pub validator_index: ValidatorIndex,

  /// The validator signature, that can be verified when constructing a
  /// `SignedDisputeStatement`.
  pub signature: ValidatorSignature,

  /// Kind of dispute statement.
  pub kind: InvalidDisputeStatementKind,
}

/// Any valid vote (backing, approval, explicit).
pub struct ValidDisputeVote {
  /// The voting validator index.
  pub validator_index: ValidatorIndex,

  /// The validator signature, that can be verified when constructing a
  /// `SignedDisputeStatement`.
  pub signature: ValidatorSignature,

  /// Kind of dispute statement.
  pub kind: ValidDisputeStatementKind,
}
}

Response:

#![allow(unused)]
fn main() {
enum DisputeResponse {
  Confirmed
}
}

Vote Recovery

Protocol: "/<genesis_hash>/<fork_id>/req_votes/1"

#![allow(unused)]
fn main() {
struct IHaveVotesRequest {
  candidate_hash: CandidateHash,
  session: SessionIndex,
  valid_votes: Bitfield,
  invalid_votes: Bitfield,
}

}

Response:

#![allow(unused)]
fn main() {
struct VotesResponse {
  /// All votes we have, but the requester was missing.
  missing: Vec<(DisputeStatement, ValidatorIndex, ValidatorSignature)>,
}
}

Starting a Dispute

A dispute is initiated once a node sends the first DisputeRequest wire message, which must contain an "invalid" vote and a "valid" vote.

The dispute distribution subsystem can get instructed to send that message out to all concerned validators by means of a DisputeDistributionMessage::SendDispute message. That message must contain an invalid vote from the local node and some valid one, e.g. a backing statement.

We include a valid vote as well, so any node regardless of whether it is synced with the chain or not or has seen backing/approval vote can see that there are conflicting votes available, hence we have a valid dispute. Nodes will still need to check whether the disputing votes are somewhat current and not some stale ones.

Participating in a Dispute

Upon receiving a DisputeRequest message, a dispute distribution will trigger the import of the received votes via the dispute coordinator (DisputeCoordinatorMessage::ImportStatements). The dispute coordinator will take care of participating in that dispute if necessary. Once it is done, the coordinator will send a DisputeDistributionMessage::SendDispute message to dispute distribution. From here, everything is the same as for starting a dispute, except that if the local node deemed the candidate valid, the SendDispute message will contain a valid vote signed by our node and will contain the initially received Invalid vote.

Note, that we rely on dispute-coordinator to check validity of a dispute for spam protection (see below).

Sending of messages

Starting and participating in a dispute are pretty similar from the perspective of dispute distribution. Once we receive a SendDispute message, we try to make sure to get the data out. We keep track of all the parachain validators that should see the message, which are all the parachain validators of the session where the dispute happened as they will want to participate in the dispute. In addition we also need to get the votes out to all authorities of the current session (which might be the same or not and may change during the dispute). Those authorities will not participate in the dispute, but need to see the statements so they can include them in blocks.

Reliability

We only consider a message transmitted, once we received a confirmation message. If not, we will keep retrying getting that message out as long as the dispute is deemed alive. To determine whether a dispute is still alive we will ask the dispute-coordinator for a list of all still active disputes via a DisputeCoordinatorMessage::ActiveDisputes message before each retry run. Once a dispute is no longer live, we will clean up the state accordingly.

Order

We assume SendDispute messages are coming in an order of importance, hence dispute-distribution will make sure to send out network messages in the same order, even on retry.

Rate Limit

For spam protection (see below), we employ an artificial rate limiting on sending out messages in order to not hit the rate limit at the receiving side, which would result in our messages getting dropped and our reputation getting reduced.

Reception

As we shall see the receiving side is mostly about handling spam and ensuring the dispute-coordinator learns about disputes as fast as possible.

Goals for the receiving side:

  1. Get new disputes to the dispute-coordinator as fast as possible, so prioritization can happen properly.
  2. Batch votes per disputes as much as possible for good import performance.
  3. Prevent malicious nodes exhausting node resources by sending lots of messages.
  4. Prevent malicious nodes from sending so many messages/(fake) disputes, preventing us from concluding good ones.
  5. Limit ability of malicious nodes of delaying the vote import due to batching logic.

Goal 1 and 2 seem to be conflicting, but an easy compromise is possible: When learning about a new dispute, we will import the vote immediately, making the dispute coordinator aware and also getting immediate feedback on the validity. Then if valid we can batch further incoming votes, with less time constraints as the dispute-coordinator already knows about the dispute.

Goal 3 and 4 are obviously very related and both can easily be solved via rate limiting as we shall see below. Rate limits should already be implemented at the substrate level, but are not at the time of writing. But even if they were, the enforced substrate limits would likely not be configurable and thus would still be to high for our needs as we can rely on the following observations:

  1. Each honest validator will only send one message (apart from duplicates on timeout) per candidate/dispute.
  2. An honest validator needs to fully recover availability and validate the candidate for casting a vote.

With these two observations, we can conclude that honest validators will usually not send messages at a high rate. We can therefore enforce conservative rate limits and thus minimize harm spamming malicious nodes can have.

Before we dive into how rate limiting solves all spam issues elegantly, let's discuss that honest behaviour further:

What about session changes? Here we might have to inform a new validator set of lots of already existing disputes at once.

With observation 1) and a rate limit that is per peer, we are still good:

Let's assume a rate limit of one message per 200ms per sender. This means 5 messages from each validator per second. 5 messages means 5 disputes! Conclusively, we will be able to conclude 5 disputes per second - no matter what malicious actors are doing. This is assuming dispute messages are sent ordered, but even if not perfectly ordered: On average it will be 5 disputes per second.

This is good enough! All those disputes are valid ones and will result in slashing and disabling of validators. Let's assume all of them conclude valid, and we disable validators only after 100 raised concluding valid disputes, we would still start disabling misbehaving validators in only 20 seconds.

One could also think that in addition participation is expected to take longer, which means on average we can import/conclude disputes faster than they are generated - regardless of dispute spam. Unfortunately this is not necessarily true: There might be parachains with very light load where recovery and validation can be accomplished very quickly - maybe faster than we can import those disputes.

This is probably an argument for not imposing a too low rate limit, although the issue is more general: Even without any rate limit, if an attacker generates disputes at a very high rate, nodes will be having trouble keeping participation up, hence the problem should be mitigated at a more fundamental layer.

For nodes that have been offline for a while, the same argument as for session changes holds, but matters even less: We assume 2/3 of nodes to be online, so even if the worst case 1/3 offline happens and they could not import votes fast enough (as argued above, they in fact can) it would not matter for consensus.

Rate Limiting

As suggested previously, rate limiting allows to mitigate all threats that come from malicious actors trying to overwhelm the system in order to get away without a slash, when it comes to dispute-distribution. In this section we will explain how in greater detail.

The idea is to open a queue with limited size for each peer. We will process incoming messages as fast as we can by doing the following:

  1. Check that the sending peer is actually a valid authority - otherwise drop message and decrease reputation/disconnect.
  2. Put message on the peer's queue, if queue is full - drop it.

Every RATE_LIMIT seconds (or rather milliseconds), we pause processing incoming requests to go a full circle and process one message from each queue. Processing means Batching as explained in the next section.

Batching

To achieve goal 2 we will batch incoming votes/messages together before passing them on as a single batch to the dispute-coordinator. To adhere to goal 1 as well, we will do the following:

  1. For an incoming message, we check whether we have an existing batch for that candidate, if not we import directly to the dispute-coordinator, as we have to assume this is concerning a new dispute.
  2. We open a batch and start collecting incoming messages for that candidate, instead of immediately forwarding.
  3. We keep collecting votes in the batch until we receive less than MIN_KEEP_BATCH_ALIVE_VOTES unique votes in the last BATCH_COLLECTING_INTERVAL. This is important to accommodate for goal 5 and also 3.
  4. We send the whole batch to the dispute-coordinator.

This together with rate limiting explained above ensures we will be able to process valid disputes: We can limit the number of simultaneous existing batches to some high value, but can be rather certain that this limit will never be reached - hence we won't drop valid disputes:

Let's assume MIN_KEEP_BATCH_ALIVE_VOTES is 10, BATCH_COLLECTING_INTERVAL is 500ms and above RATE_LIMIT is 100ms. 1/3 of validators are malicious, so for 1000 this means around 330 malicious actors worst case.

All those actors can send a message every 100ms, that is 10 per second. This means at the begining of an attack they can open up around 3300 batches. Each containing two votes. So memory usage is still negligible. In reality it is even less, as we also demand 10 new votes to trickle in per batch in order to keep it alive, every 500ms. Hence for the first second, each batch requires 20 votes each. Each message is 2 votes, so this means 10 messages per batch. Hence to keep those batches alive 10 attackers are needed for each batch. This reduces the number of opened batches by a factor of 10: So we only have 330 batches in 1 second - each containing 20 votes.

The next second: In order to further grow memory usage, attackers have to maintain 10 messages per batch and second. Number of batches equals the number of attackers, each has 10 messages per second, all are needed to maintain the batches in memory. Therefore we have a hard cap of around 330 (number of malicious nodes) open batches. Each can be filled with number of malicious actor's votes. So 330 batches with each 330 votes: Let's assume approximately 100 bytes per signature/vote. This results in a worst case memory usage of 330 * 330

  • 100 ~= 10 MiB.

For 10_000 validators, we are already in the Gigabyte range, which means that with a validator set that large we might want to be more strict with the rate limit or require a larger rate of incoming votes per batch to keep them alive.

For a thousand validators a limit on batches of around 1000 should never be reached in practice. Hence due to rate limiting we have a very good chance to not ever having to drop a potential valid dispute due to some resource limit.

Further safe guards are possible: The dispute-coordinator actually confirms/denies imports. So once we receive a denial by the dispute-coordinator for the initial imported votes, we can opt into flushing the batch immediately and importing the votes. This swaps memory usage for more CPU usage, but if that import is deemed invalid again we can immediately decrease the reputation of the sending peers, so this should be a net win. For the time being we punt on this for simplicity.

Instead of filling batches to maximize memory usage, attackers could also try to overwhelm the dispute coordinator by only sending votes for new candidates all the time. This attack vector is mitigated also by above rate limit and decreasing the peer's reputation on denial of the invalid imports by the coordinator.

Node Startup

Nothing special happens on node startup. We expect the dispute-coordinator to inform us about any ongoing disputes via SendDispute messages.

Backing and Approval Votes

Backing and approval votes get imported when they arrive/are created via the dispute coordinator by corresponding subsystems.

We assume that under normal operation each node will be aware of backing and approval votes and optimize for that case. Nevertheless we want disputes to conclude fast and reliable, therefore if a node is not aware of backing/approval votes it can request the missing votes from the node that informed it about the dispute (see Resiliency

Resiliency

The above protocol should be sufficient for most cases, but there are certain cases we also want to have covered:

  • Non validator nodes might be interested in ongoing voting, even before it is recorded on chain.
  • Nodes might have missed votes, especially backing or approval votes. Recovering them from chain is difficult and expensive, due to runtime upgrades and untyped extrinsics.
  • More importantly, on era changes the new authority set, from the perspective of approval-voting have no need to see "old" approval votes, hence they might not see them, can therefore not import them into the dispute coordinator and therefore no authority will put them on chain.

To cover those cases, we introduce a second request/response protocol, which can be handled on a lower priority basis as the one above. It consists of the request/response messages as described in the [protocol section][#vote-recovery].

Nodes may send those requests to validators, if they feel they are missing votes. E.g. after some timeout, if no majority was reached yet in their point of view or if they are not aware of any backing/approval votes for a received disputed candidate.

The receiver of a IHaveVotesRequest message will do the following:

  1. See if the sender is missing votes we are aware of - if so, respond with those votes.
  2. Check whether the sender knows about any votes, we don't know about and if so send a IHaveVotesRequest request back, with our knowledge.
  3. Record the peer's knowledge.

When to send IHaveVotesRequest messages:

  1. Whenever we are asked to do so via DisputeDistributionMessage::FetchMissingVotes.
  2. Approximately once per block to some random validator as long as the dispute is active.

Spam considerations: Nodes want to accept those messages once per validator and per slot. They are free to drop more frequent requests or requests for stale data. Requests coming from non validator nodes, can be handled on a best effort basis.

Considerations

Dispute distribution is critical. We should keep track of available validator connections and issue warnings if we are not connected to a majority of validators. We should also keep track of failed sending attempts and log warnings accordingly. As disputes are rare and TCP is a reliable protocol, probably each failed attempt should trigger a warning in logs and also logged into some Prometheus metric.

Disputes for non available candidates

If deemed necessary we can later on also support disputes for non available candidates, but disputes for those cases have totally different requirements.

First of all such disputes are not time critical. We just want to have some offender slashed at some point, but we have no risk of finalizing any bad data.

Second, as we won't have availability for such data, the node that initiated the dispute will be responsible for providing the disputed data initially. Then nodes which did the check already are also providers of the data, hence distributing load and making prevention of the dispute from concluding harder and harder over time. Assuming an attacker can not DoS a node forever, the dispute will succeed eventually, which is all that matters. And again, even if an attacker managed to prevent such a dispute from happening somehow, there is no real harm done: There was no serious attack to begin with.

Utility Subsystems

The utility subsystems are an assortment which don't have a natural home in another subsystem collection.

Availability Store

This is a utility subsystem responsible for keeping available certain data and pruning that data.

The two data types:

  • Full PoV blocks of candidates we have validated
  • Availability chunks of candidates that were backed and noted available on-chain.

For each of these data we have pruning rules that determine how long we need to keep that data available.

PoV hypothetically only need to be kept around until the block where the data was made fully available is finalized. However, disputes can revert finality, so we need to be a bit more conservative and we add a delay. We should keep the PoV until a block that finalized availability of it has been finalized for 1 day + 1 hour.

Availability chunks need to be kept available until the dispute period for the corresponding candidate has ended. We can accomplish this by using the same criterion as the above. This gives us a pruning condition of the block finalizing availability of the chunk being final for 1 day + 1 hour.

There is also the case where a validator commits to make a PoV available, but the corresponding candidate is never backed. In this case, we keep the PoV available for 1 hour.

There may be multiple competing blocks all ending the availability phase for a particular candidate. Until finality, it will be unclear which of those is actually the canonical chain, so the pruning records for PoVs and Availability chunks should keep track of all such blocks.

Lifetime of the block data and chunks in storage

%3Block data FSMstStoredincIncludedst->incBlockincludedprnPrunedst->prnStored blocktimed outinc->stCompeting blocksfinalizedfinFinalizedinc->finBlockfinalizedfin->prnBlock keep time(1 day + 1 hour) elapsed

Database Schema

We use an underlying Key-Value database where we assume we have the following operations available:

  • write(key, value)
  • read(key) -> Option<value>
  • iter_with_prefix(prefix) -> Iterator<(key, value)> - gives all keys and values in lexicographical order where the key starts with prefix.

We use this database to encode the following schema:

#![allow(unused)]
fn main() {
("available", CandidateHash) -> Option<AvailableData>
("chunk", CandidateHash, u32) -> Option<ErasureChunk>
("meta", CandidateHash) -> Option<CandidateMeta>

("unfinalized", BlockNumber, BlockHash, CandidateHash) -> Option<()>
("prune_by_time", Timestamp, CandidateHash) -> Option<()>
}

Timestamps are the wall-clock seconds since Unix epoch. Timestamps and block numbers are both encoded as big-endian so lexicographic order is ascending.

The meta information that we track per-candidate is defined as the CandidateMeta struct

#![allow(unused)]
fn main() {
struct CandidateMeta {
  state: State,
  data_available: bool,
  chunks_stored: Bitfield,
}

enum State {
  /// Candidate data was first observed at the given time but is not available in any block.
  Unavailable(Timestamp),
  /// The candidate was first observed at the given time and was included in the given list of unfinalized blocks, which may be
  /// empty. The timestamp here is not used for pruning. Either one of these blocks will be finalized or the state will regress to
  /// `State::Unavailable`, in which case the same timestamp will be reused.
  Unfinalized(Timestamp, Vec<(BlockNumber, BlockHash)>),
  /// Candidate data has appeared in a finalized block and did so at the given time.
  Finalized(Timestamp)
}
}

We maintain the invariant that if a candidate has a meta entry, its available data exists on disk if data_available is true. All chunks mentioned in the meta entry are available.

Additionally, there is exactly one prune_by_time entry which holds the candidate hash unless the state is Unfinalized. There may be zero, one, or many "unfinalized" keys with the given candidate, and this will correspond to the state of the meta entry.

Protocol

Input: AvailabilityStoreMessage

Output:

Functionality

For each head in the activated list:

  • Load all ancestors of the head back to the finalized block so we don't miss anything if import notifications are missed. If a StoreChunk message is received for a candidate which has no entry, then we will prematurely lose the data.
  • Note any new candidates backed in the head. Update the CandidateMeta for each. If the CandidateMeta does not exist, create it as Unavailable with the current timestamp. Register a "prune_by_time" entry based on the current timestamp + 1 hour.
  • Note any new candidate included in the head. Update the CandidateMeta for each, performing a transition from Unavailable to Unfinalized if necessary. That includes removing the "prune_by_time" entry. Add the head hash and number to the state, if unfinalized. Add an "unfinalized" entry for the block and candidate.
  • The CandidateEvent runtime API can be used for this purpose.

On OverseerSignal::BlockFinalized(finalized) events:

  • for each key in iter_with_prefix("unfinalized")

    • Stop if the key is beyond ("unfinalized, finalized)
    • For each block number f that we encounter, load the finalized hash for that block.
      • The state of each CandidateMeta we encounter here must be Unfinalized, since we loaded the candidate from an "unfinalized" key.
      • For each candidate that we encounter under f and the finalized block hash,
        • Update the CandidateMeta to have State::Finalized. Remove all "unfinalized" entries from the old Unfinalized state.
        • Register a "prune_by_time" entry for the candidate based on the current time + 1 day + 1 hour.
      • For each candidate that we encounter under f which is not under the finalized block hash,
        • Remove all entries under f in the Unfinalized state.
        • If the CandidateMeta has state Unfinalized with an empty list of blocks, downgrade to Unavailable and re-schedule pruning under the timestamp + 1 hour. We do not prune here as the candidate still may be included in a descendant of the finalized chain.
      • Remove all "unfinalized" keys under f.
  • Update last_finalized = finalized.

    This is roughly O(n * m) where n is the number of blocks finalized since the last update, and m is the number of parachains.

On QueryAvailableData message:

  • Query ("available", candidate_hash)

    This is O(n) in the size of the data, which may be large.

On QueryDataAvailability message:

  • Query whether ("meta", candidate_hash) exists and data_available == true.

    This is O(n) in the size of the metadata which is small.

On QueryChunk message:

  • Query ("chunk", candidate_hash, index)

    This is O(n) in the size of the data, which may be large.

On QueryAllChunks message:

  • Query ("meta", candidate_hash). If None, send an empty response and return.
  • For all 1 bits in the chunks_stored, query ("chunk", candidate_hash, index). Ignore but warn on errors, and return a vector of all loaded chunks.

On QueryChunkAvailability message:

  • Query whether ("meta", candidate_hash) exists and the bit at index is set.

    This is O(n) in the size of the metadata which is small.

On StoreChunk message:

  • If there is a CandidateMeta under the candidate hash, set the bit of the erasure-chunk in the chunks_stored bitfield to 1. If it was not 1 already, write the chunk under ("chunk", candidate_hash, chunk_index).

    This is O(n) in the size of the chunk.

On StoreAvailableData message:

  • Compute the erasure root of the available data and compare it with expected_erasure_root. Return StoreAvailableDataError::InvalidErasureRoot on mismatch.

  • If there is no CandidateMeta under the candidate hash, create it with State::Unavailable(now). Load the CandidateMeta otherwise.

  • Store data under ("available", candidate_hash) and set data_available to true.

  • Store each chunk under ("chunk", candidate_hash, index) and set every bit in chunks_stored to 1.

    This is O(n) in the size of the data as the aggregate size of the chunks is proportional to the data.

Every 5 minutes, run a pruning routine:

  • for each key in iter_with_prefix("prune_by_time"):

    • If the key is beyond ("prune_by_time", now), return.
    • Remove the key.
    • Extract candidate_hash from the key.
    • Load and remove the ("meta", candidate_hash)
    • For each erasure chunk bit set, remove ("chunk", candidate_hash, bit_index).
    • If data_available, remove ("available", candidate_hash) This is O(n * m) in the amount of candidates and average size of the data stored. This is probably the most expensive operation but does not need to be run very often.

Basic scenarios to test

Basically we need to test the correctness of data flow through state FSMs described earlier. These tests obviously assume that some mocking of time is happening.

  • Stored data that is never included pruned in necessary timeout

    • A block (and/or a chunk) is added to the store.
    • We never note that the respective candidate is included.
    • Until a defined timeout the data in question is available.
    • After this timeout the data is no longer available.
  • Stored data is kept until we are certain it is finalized.

    • A block (and/or a chunk) is added to the store.
    • It is available.
    • Before the inclusion timeout expires notify storage that the candidate was included.
    • The data is still available.
    • Wait for an absurd amount of time (longer than 1 day).
    • Check that the data is still available.
    • Send finality notification about the block in question.
    • Wait for some time below finalized data timeout.
    • The data is still available.
    • Wait until the data should have been pruned.
    • The data is no longer available.
  • Fork-awareness of the relay chain is taken into account

    • Block B1 is added to the store.
    • Block B2 is added to the store.
    • Notify the subsystem that both B1 and B2 were included in different leafs of relay chain.
    • Notify the subsystem that the leaf with B1 was finalized.
    • Leaf with B2 is never finalized.
    • Leaf with B2 is pruned and its data is no longer available.
    • Wait until the finalized data of B1 should have been pruned.
    • B1 is no longer available.

Candidate Validation

This subsystem is responsible for handling candidate validation requests. It is a simple request/response server.

A variety of subsystems want to know if a parachain block candidate is valid. None of them care about the detailed mechanics of how a candidate gets validated, just the results. This subsystem handles those details.

Protocol

Input: CandidateValidationMessage

Output: Validation result via the provided response side-channel.

Functionality

This subsystem groups the requests it handles in two categories: candidate validation and PVF pre-checking.

The first category can be further subdivided in two request types: one which draws out validation data from the state, and another which accepts all validation data exhaustively. Validation returns three possible outcomes on the response channel: the candidate is valid, the candidate is invalid, or an internal error occurred.

Parachain candidates are validated against their validation function: A piece of Wasm code that describes the state-transition of the parachain. Validation function execution is not metered. This means that an execution which is an infinite loop or simply takes too long must be forcibly exited by some other means. For this reason, we recommend dispatching candidate validation to be done on subprocesses which can be killed if they time-out.

Upon receiving a validation request, the first thing the candidate validation subsystem should do is make sure it has all the necessary parameters to the validation function. These are:

The second category is for PVF pre-checking. This is primarly used by the PVF pre-checker subsystem.

Determining Parameters

For a CandidateValidationMessage::ValidateFromExhaustive, these parameters are exhaustively provided.

For a CandidateValidationMessage::ValidateFromChainState, some more work needs to be done. Due to the uncertainty of Availability Cores (implemented in the Scheduler module of the runtime), a candidate at a particular relay-parent and for a particular para may have two different valid validation-data to be executed under depending on what is assumed to happen if the para is occupying a core at the onset of the new block. This is encoded as an OccupiedCoreAssumption in the runtime API.

The way that we can determine which assumption the candidate is meant to be executed under is simply to do an exhaustive check of both possibilities based on the state of the relay-parent. First we fetch the validation data under the assumption that the block occupying becomes available. If the validation_data_hash of the CandidateDescriptor matches this validation data, we use that. Otherwise, if the validation_data_hash matches the validation data fetched under the TimedOut assumption, we use that. Otherwise, we return a ValidationResult::Invalid response and conclude.

Then, we can fetch the validation code from the runtime based on which type of candidate this is. This gives us all the parameters. The descriptor and PoV come from the request itself, and the other parameters have been derived from the state.

TODO: This would be a great place for caching to avoid making lots of runtime requests. That would need a job, though.

Execution of the Parachain Wasm

Once we have all parameters, we can spin up a background task to perform the validation in a way that doesn't hold up the entire event loop. Before invoking the validation function itself, this should first do some basic checks:

  • The collator signature is valid
  • The PoV provided matches the pov_hash field of the descriptor

For more details please see PVF Host and Workers.

Checking Validation Outputs

If we can assume the presence of the relay-chain state (that is, during processing CandidateValidationMessage::ValidateFromChainState) we can run all the checks that the relay-chain would run at the inclusion time thus confirming that the candidate will be accepted.

PVF Host and Workers

The PVF host is responsible for handling requests to prepare and execute PVF code blobs, which it sends to PVF workers running in their own child processes.

This system has two high-levels goals that we will touch on here: determinism and security.

Determinism

One high-level goal is to make PVF operations as deterministic as possible, to reduce the rate of disputes. Disputes can happen due to e.g. a job timing out on one machine, but not another. While we do not have full determinism, there are some dispute reduction mechanisms in place right now.

Retrying execution requests

If the execution request fails during preparation, we will retry if it is possible that the preparation error was transient (e.g. if the error was a panic or time out). We will only retry preparation if another request comes in after 15 minutes, to ensure any potential transient conditions had time to be resolved. We will retry up to 5 times.

If the actual execution of the artifact fails, we will retry once if it was a possibly transient error, to allow the conditions that led to the error to hopefully resolve. We use a more brief delay here (1 second as opposed to 15 minutes for preparation (see above)), because a successful execution must happen in a short amount of time.

We currently know of the following specific cases that will lead to a retried execution request:

  1. OOM: The host might have been temporarily low on memory due to other processes running on the same machine. NOTE: This case will lead to voting against the candidate (and possibly a dispute) if the retry is still not successful.
  2. Artifact missing: The prepared artifact might have been deleted due to operator error or some bug in the system.
  3. Panic: The worker thread panicked for some indeterminate reason, which may or may not be independent of the candidate or PVF.

Preparation timeouts

We use timeouts for both preparation and execution jobs to limit the amount of time they can take. As the time for a job can vary depending on the machine and load on the machine, this can potentially lead to disputes where some validators successfuly execute a PVF and others don't.

One dispute mitigation we have in place is a more lenient timeout for preparation during execution than during pre-checking. The rationale is that the PVF has already passed pre-checking, so we know it should be valid, and we allow it to take longer than expected, as this is likely due to an issue with the machine and not the PVF.

CPU clock timeouts

Another timeout-related mitigation we employ is to measure the time taken by jobs using CPU time, rather than wall clock time. This is because the CPU time of a process is less variable under different system conditions. When the overall system is under heavy load, the wall clock time of a job is affected more than the CPU time.

Internal errors

In general, for errors not raising a dispute we have to be very careful. This is only sound, if we either:

  1. Ruled out that error in pre-checking. If something is not checked in pre-checking, even if independent of the candidate and PVF, we must raise a dispute.
  2. We are 100% confident that it is a hardware/local issue: Like corrupted file, etc.

Reasoning: Otherwise it would be possible to register a PVF where candidates can not be checked, but we don't get a dispute - so nobody gets punished. Second, we end up with a finality stall that is not going to resolve!

There are some error conditions where we can't be sure whether the candidate is really invalid or some internal glitch occurred, e.g. panics. Whenever we are unsure, we can never treat an error as internal as we would abstain from voting. So we will first retry the candidate, and if the issue persists we are forced to vote invalid.

Security

With on-demand parachains, it is much easier to submit PVFs to the chain for preparation and execution. This makes it easier for erroneous disputes and slashing to occur, whether intentional (as a result of a malicious attacker) or not (a bug or operator error occurred).

Therefore, another goal of ours is to harden our security around PVFs, in order to protect the economic interests of validators and increase overall confidence in the system.

Possible attacks / threat model

Webassembly is already sandboxed, but there have already been reported multiple CVEs enabling remote code execution. See e.g. these two advisories from Mar 2023 and Jul 2022.

So what are we actually worried about? Things that come to mind:

  1. Consensus faults - If an attacker can get some source of randomness they could vote against with 50% chance and cause unresolvable disputes.
  2. Targeted slashes - An attacker can target certain validators (e.g. some validators running on vulnerable hardware) and make them vote invalid and get them slashed.
  3. Mass slashes - With some source of randomness they can do an untargeted attack. I.e. a baddie can do significant economic damage by voting against with 1/3 chance, without even stealing keys or completely replacing the binary.
  4. Stealing keys - That would be pretty bad. Should not be possible with sandboxing. We should at least not allow filesystem-access or network access.
  5. Taking control over the validator. E.g. replacing the polkadot binary with a polkadot-evil binary. Should again not be possible with the above sandboxing in place.
  6. Intercepting and manipulating packages - Effect very similar to the above, hard to do without also being able to do 4 or 5.

Restricting file-system access

A basic security mechanism is to make sure that any thread directly interfacing with untrusted code does not have access to the file-system. This provides some protection against attackers accessing sensitive data or modifying data on the host machine.

Clearing env vars

We clear environment variables before handling untrusted code, because why give attackers potentially sensitive data unnecessarily? And even if everything else is locked down, env vars can potentially provide a source of randomness (see point 1, "Consensus faults" above).

Provisioner

Relay chain block authorship authority is governed by BABE and is beyond the scope of the Overseer and the rest of the subsystems. That said, ultimately the block author needs to select a set of backable parachain candidates and other consensus data, and assemble a block from them. This subsystem is responsible for providing the necessary data to all potential block authors.

Provisionable Data

There are several distinct types of provisionable data, but they share this property in common: all should eventually be included in a relay chain block.

Backed Candidates

The block author can choose 0 or 1 backed parachain candidates per parachain; the only constraint is that each backable candidate has the appropriate relay parent. However, the choice of a backed candidate must be the block author's. The provisioner subsystem is how those block authors make this choice in practice.

Signed Bitfields

Signed bitfields are attestations from a particular validator about which candidates it believes are available. Those will only be provided on fresh leaves.

Misbehavior Reports

Misbehavior reports are self-contained proofs of misbehavior by a validator or group of validators. For example, it is very easy to verify a double-voting misbehavior report: the report contains two votes signed by the same key, advocating different outcomes. Concretely, misbehavior reports become inherents which cause dots to be slashed.

Note that there is no mechanism in place which forces a block author to include a misbehavior report which it doesn't like, for example if it would be slashed by such a report. The chain's defense against this is to have a relatively long slash period, such that it's likely to encounter an honest author before the slash period expires.

Dispute Inherent

The dispute inherent is similar to a misbehavior report in that it is an attestation of misbehavior on the part of a validator or group of validators. Unlike a misbehavior report, it is not self-contained: resolution requires coordinated action by several validators. The canonical example of a dispute inherent involves an approval checker discovering that a set of validators has improperly approved an invalid parachain block: resolving this requires the entire validator set to re-validate the block, so that the minority can be slashed.

Dispute resolution is complex and is explained in substantially more detail here.

Protocol

The subsystem should maintain a set of handles to Block Authorship Provisioning iterations that are currently live.

On Overseer Signal

  • ActiveLeavesUpdate:
    • For each activated head:
      • spawn a Block Authorship Provisioning iteration with the given relay parent, storing a bidirectional channel with that iteration.
    • For each deactivated head:
      • terminate the Block Authorship Provisioning iteration for the given relay parent, if any.
  • Conclude: Forward Conclude to all iterations, waiting a small amount of time for them to join, and then hard-exiting.

On ProvisionerMessage

Forward the message to the appropriate Block Authorship Provisioning iteration, or discard if no appropriate iteration is currently active.

Per Provisioning Iteration

Input: ProvisionerMessage. Backed candidates come from the Candidate Backing subsystem, signed bitfields come from the Bitfield Distribution subsystem, and disputes come from the Disputes Subsystem. Misbehavior reports are currently sent from the Candidate Backing subsystem and contain the following misbehaviors:

  1. Misbehavior::ValidityDoubleVote
  2. Misbehavior::MultipleCandidates
  3. Misbehavior::UnauthorizedStatement
  4. Misbehavior::DoubleSign

But we choose not to punish these forms of misbehavior for the time being. Risks from misbehavior are sufficiently mitigated at the protocol level via reputation changes. Punitive actions here may become desirable enough to dedicate time to in the future.

At initialization, this subsystem has no outputs.

Block authors request the inherent data they should use for constructing the inherent in the block which contains parachain execution information.

Block Production

When a validator is selected by BABE to author a block, it becomes a block producer. The provisioner is the subsystem best suited to choosing which specific backed candidates and availability bitfields should be assembled into the block. To engage this functionality, a ProvisionerMessage::RequestInherentData is sent; the response is a ParaInherentData. Each relay chain block backs at most one backable parachain block candidate per parachain. Additionally no further block candidate can be backed until the previous one either gets declared available or expired. If bitfields indicate that candidate A, predecessor of B, should be declared available, then B can be backed in the same relay block. Appropriate bitfields, as outlined in the section on bitfield selection, and any dispute statements should be attached as well.

Bitfield Selection

Our goal with respect to bitfields is simple: maximize availability. However, it's not quite as simple as always including all bitfields; there are constraints which still need to be met:

  • not more than one bitfield per validator
  • each 1 bit must correspond to an occupied core

Beyond that, a semi-arbitrary selection policy is fine. In order to meet the goal of maximizing availability, a heuristic of picking the bitfield with the greatest number of 1 bits set in the event of conflict is useful.

Dispute Statement Selection

This is the point at which the block author provides further votes to active disputes or initiates new disputes in the runtime state.

The block-authoring logic of the runtime has an extra step between handling the inherent-data and producing the actual inherent call, which we assume performs the work of filtering out disputes which are not relevant to the on-chain state. Backing votes are always kept in the dispute statement set. This ensures we punish the maximum number of misbehaving backers.

To select disputes:

  • Issue a DisputeCoordinatorMessage::RecentDisputes message and wait for the response. This is a set of all disputes in recent sessions which we are aware of.

Determining Bitfield Availability

An occupied core has a CoreAvailability bitfield. We also have a list of SignedAvailabilityBitfields. We need to determine from these whether or not a core at a particular index has become available.

The key insight required is that CoreAvailability is transverse to the SignedAvailabilityBitfields: if we conceptualize the list of bitfields as many rows, each bit of which is its own column, then CoreAvailability for a given core index is the vertical slice of bits in the set at that index.

To compute bitfield availability, then:

  • Start with a copy of OccupiedCore.availability
  • For each bitfield in the list of SignedAvailabilityBitfields:
    • Get the bitfield's validator_index
    • Update the availability. Conceptually, assuming bit vectors: availability[validator_index] |= bitfield[core_idx]
  • Availability has a 2/3 threshold. Therefore: 3 * availability.count_ones() >= 2 * availability.len()

Candidate Selection: Prospective Parachains Mode

The state of the provisioner PerRelayParent tracks an important setting, ProspectiveParachainsMode. This setting determines which backable candidate selection method the provisioner uses.

ProspectiveParachainsMode::Disabled - The provisioner uses its own internal legacy candidate selection. ProspectiveParachainsMode::Enabled - The provisioner requests that prospective parachains provide selected candidates.

Candidates selected with ProspectiveParachainsMode::Enabled are able to benefit from the increased block production time asynchronous backing allows. For this reason all Polkadot protocol networks will eventually use prospective parachains candidate selection. Then legacy candidate selection will be removed as obsolete.

Prospective Parachains Candidate Selection

The goal of candidate selection is to determine which cores are free, and then to the degree possible, pick a candidate appropriate to each free core. In prospective parachains candidate selection the provisioner handles the former process while prospective parachains handles the latter.

To select backable candidates:

  • Get the list of core states from the runtime API
  • For each core state:
    • On CoreState::Free
      • The core is unscheduled and doesn’t need to be provisioned with a candidate
    • On CoreState::Scheduled
      • The core is unoccupied and scheduled to accept a backed block for a particular para_id.
      • The provisioner requests a backable candidate from prospective parachains with the desired relay parent, the core’s scheduled para_id, and an empty required path.
    • On CoreState::Occupied
      • The availability core is occupied by a parachain block candidate pending availability. A further candidate need not be provided by the provisioner unless the core will be vacated this block. This is the case when either bitfields indicate the current core occupant has been made available or a timeout is reached.
      • If bitfields_indicate_availability
        • If Some(scheduled_core) = occupied_core.next_up_on_available, the core will be vacated and in need of a provisioned candidate. The provisioner requests a backable candidate from prospective parachains with the core’s scheduled para_id and a required path with one entry. This entry corresponds to the parablock candidate previously occupying this core, which was made available and can be built upon even though it hasn’t been seen as included in a relay chain block yet. See the Required Path section below for more detail.
        • If occupied_core.next_up_on_available is None, then the core being vacated is unscheduled and doesn’t need to be provisioned with a candidate.
      • Else-if occupied_core.time_out_at == block_number
        • If Some(scheduled_core) = occupied_core.next_up_on_timeout, the core will be vacated and in need of a provisioned candidate. A candidate is requested in exactly the same way as with CoreState::Scheduled.
        • Else the core being vacated is unscheduled and doesn’t need to be provisioned with a candidate The end result of this process is a vector of CandidateHashs, sorted in order of their core index.

Required Path

Required path is a parameter for ProspectiveParachainsMessage::GetBackableCandidate, which the provisioner sends in candidate selection.

An empty required path indicates that the requested candidate should be a direct child of the most recently included parablock for the given para_id as of the given relay parent.

In contrast, a required path with one or more entries prompts prospective parachains to step forward through its fragment tree for the given para_id and relay parent until the desired parablock is reached. We then select a direct child of that parablock to pass to the provisioner.

The parablocks making up a required path do not need to have been previously seen as included in relay chain blocks. Thus the ability to provision backable candidates based on a required path effectively decouples backing from inclusion.

Legacy Candidate Selection

Legacy candidate selection takes place in the provisioner. Thus the provisioner needs to keep an up to date record of all backed_candidates PerRelayParent to pick from.

The goal of candidate selection is to determine which cores are free, and then to the degree possible, pick a candidate appropriate to each free core.

To determine availability:

  • Get the list of core states from the runtime API
  • For each core state:
    • On CoreState::Scheduled, then we can make an OccupiedCoreAssumption::Free.
    • On CoreState::Occupied, then we may be able to make an assumption:
      • If the bitfields indicate availability and there is a scheduled next_up_on_available, then we can make an OccupiedCoreAssumption::Included.
      • If the bitfields do not indicate availability, and there is a scheduled next_up_on_time_out, and occupied_core.time_out_at == block_number_under_production, then we can make an OccupiedCoreAssumption::TimedOut.
    • If we did not make an OccupiedCoreAssumption, then continue on to the next core.
    • Now compute the core's validation_data_hash: get the PersistedValidationData from the runtime, given the known ParaId and OccupiedCoreAssumption;
    • Find an appropriate candidate for the core.
      • There are two constraints: backed_candidate.candidate.descriptor.para_id == scheduled_core.para_id && candidate.candidate.descriptor.validation_data_hash == computed_validation_data_hash.
      • In the event that more than one candidate meets the constraints, selection between the candidates is arbitrary. However, not more than one candidate can be selected per core.

The end result of this process is a vector of CandidateHashs, sorted in order of their core index.

Retrieving Full BackedCandidates for Selected Hashes

Legacy candidate selection and prospective parachains candidate selection both leave us with a vector of CandidateHashs. These are passed to the backing subsystem with CandidateBackingMessage::GetBackedCandidates.

The response is a vector of BackedCandidates, sorted in order of their core index and ready to be provisioned to block authoring. The candidate selection and retrieval process should select at maximum one candidate which upgrades the runtime validation code.

Glossary

  • Relay-parent:
    • A particular relay-chain block which serves as an anchor and reference point for processes and data which depend on relay-chain state.
  • Active Leaf:
    • A relay chain block which is the head of an active fork of the relay chain.
    • Block authorship provisioning jobs are spawned per active leaf and concluded for any leaves which become inactive.
  • Candidate Selection:
    • The process by which the provisioner selects backable parachain block candidates to pass to block authoring.
    • Two versions, prospective parachains candidate selection and legacy candidate selection. See their respective protocol sections for details.
  • Availability Core:
    • Often referred to simply as "cores", availability cores are an abstraction used for resource management. For the provisioner, availability cores are most relevant in that core states determine which para_ids to provision backable candidates for.
    • For more on availability cores see Scheduler Module: Availability Cores
  • Availability Bitfield:
    • Often referred to simply as a "bitfield", an availability bitfield represents the view of parablock candidate availability from a particular validator's perspective. Each bit in the bitfield corresponds to a single availability core.
    • For more on availability bitfields see availability
  • Backable vs. Backed:
    • Note that we sometimes use "backed" to refer to candidates that are "backable", but not yet backed on chain.
    • Backable means that a quorum of the candidate's assigned backing group have provided signed affirming statements.

Network Bridge

One of the main features of the overseer/subsystem duality is to avoid shared ownership of resources and to communicate via message-passing. However, implementing each networking subsystem as its own network protocol brings a fair share of challenges.

The most notable challenge is coordinating and eliminating race conditions of peer connection and disconnection events. If we have many network protocols that peers are supposed to be connected on, it is difficult to enforce that a peer is indeed connected on all of them or the order in which those protocols receive notifications that peers have connected. This becomes especially difficult when attempting to share peer state across protocols. All of the Parachain-Host's gossip protocols eliminate DoS with a data-dependency on current chain heads. However, it is inefficient and confusing to implement the logic for tracking our current chain heads as well as our peers' on each of those subsystems. Having one subsystem for tracking this shared state and distributing it to the others is an improvement in architecture and efficiency.

One other piece of shared state to track is peer reputation. When peers are found to have provided value or cost, we adjust their reputation accordingly.

So in short, this Subsystem acts as a bridge between an actual network component and a subsystem's protocol. The implementation of the underlying network component is beyond the scope of this module. We make certain assumptions about the network component:

  • The network allows registering of protocols and multiple versions of each protocol.
  • The network handles version negotiation of protocols with peers and only connects the peer on the highest version of the protocol.
  • Each protocol has its own peer-set, although there may be some overlap.
  • The network provides peer-set management utilities for discovering the peer-IDs of validators and a means of dialing peers with given IDs.

The network bridge makes use of the peer-set feature, but is not generic over peer-set. Instead, it exposes two peer-sets that event producers can attach to: Validation and Collation. More information can be found on the documentation of the NetworkBridgeMessage.

Protocol

Input: NetworkBridgeMessage

Output:

Functionality

This network bridge sends messages of these types over the network.

#![allow(unused)]
fn main() {
enum WireMessage<M> {
	ProtocolMessage(M),
	ViewUpdate(View),
}
}

and instantiates this type twice, once using the ValidationProtocolV1 message type, and once with the CollationProtocolV1 message type.

#![allow(unused)]
fn main() {
type ValidationV1Message = WireMessage<ValidationProtocolV1>;
type CollationV1Message = WireMessage<CollationProtocolV1>;
}

Startup

On startup, we register two protocols with the underlying network utility. One for validation and one for collation. We register only version 1 of each of these protocols.

Main Loop

The bulk of the work done by this subsystem is in responding to network events, signals from the overseer, and messages from other subsystems.

Each network event is associated with a particular peer-set.

Overseer Signal: ActiveLeavesUpdate

The activated and deactivated lists determine the evolution of our local view over time. A ProtocolMessage::ViewUpdate is issued to each connected peer on each peer-set, and a NetworkBridgeEvent::OurViewChange is issued to each event handler for each protocol.

We only send view updates if the node has indicated that it has finished major blockchain synchronization.

If we are connected to the same peer on both peer-sets, we will send the peer two view updates as a result.

Overseer Signal: BlockFinalized

We update our view's finalized_number to the provided one and delay ProtocolMessage::ViewUpdate and NetworkBridgeEvent::OurViewChange till the next ActiveLeavesUpdate.

Network Event: PeerConnected

Issue a NetworkBridgeEvent::PeerConnected for each Event Handler of the peer-set and negotiated protocol version of the peer. Also issue a NetworkBridgeEvent::PeerViewChange and send the peer our current view, but only if the node has indicated that it has finished major blockchain synchronization. Otherwise, we only send the peer an empty view.

Network Event: PeerDisconnected

Issue a NetworkBridgeEvent::PeerDisconnected for each Event Handler of the peer-set and negotiated protocol version of the peer.

Network Event: ProtocolMessage

Map the message onto the corresponding Event Handler based on the peer-set this message was received on and dispatch via overseer.

Network Event: ViewUpdate

  • Check that the new view is valid and note it as the most recent view update of the peer on this peer-set.
  • Map a NetworkBridgeEvent::PeerViewChange onto the corresponding Event Handler based on the peer-set this message was received on and dispatch via overseer.

ReportPeer

  • Adjust peer reputation according to cost or benefit provided

DisconnectPeer

  • Disconnect the peer from the peer-set requested, if connected.

SendValidationMessage / SendValidationMessages

  • Issue a corresponding ProtocolMessage to each listed peer on the validation peer-set.

SendCollationMessage / SendCollationMessages

  • Issue a corresponding ProtocolMessage to each listed peer on the collation peer-set.

ConnectToValidators

  • Determine the DHT keys to use for each validator based on the relay-chain state and Runtime API.
  • Recover the Peer IDs of the validators from the DHT. There may be more than one peer ID per validator.
  • Send all (ValidatorId, PeerId) pairs on the response channel.
  • Feed all Peer IDs to peer set manager the underlying network provides.

NewGossipTopology

  • Map all AuthorityDiscoveryIds to PeerIds and issue a corresponding NetworkBridgeUpdate to all validation subsystems.

Event Handlers

Network bridge event handlers are the intended recipients of particular network protocol messages. These are each a variant of a message to be sent via the overseer.

Validation V1

  • ApprovalDistributionV1Message -> ApprovalDistributionMessage::NetworkBridgeUpdate
  • BitfieldDistributionV1Message -> BitfieldDistributionMessage::NetworkBridgeUpdate
  • StatementDistributionV1Message -> StatementDistributionMessage::NetworkBridgeUpdate

Collation V1

  • CollatorProtocolV1Message -> CollatorProtocolMessage::NetworkBridgeUpdate

Gossip Support

The Gossip Support Subsystem is responsible for keeping track of session changes and issuing a connection request to all validators in the next, current and a few past sessions if we are a validator in these sessions. The request will add all validators to a reserved PeerSet, meaning we will not reject a connection request from any validator in that set.

In addition to that, it creates a gossip overlay topology per session which limits the amount of messages sent and received to be an order of sqrt of the validators. Our neighbors in this graph will be forwarded to the network bridge with the NetworkBridgeMessage::NewGossipTopology message.

See https://github.com/paritytech/polkadot/issues/3239 for more details.

The gossip topology is used by parachain distribution subsystems, such as Bitfield Distribution, (small) Statement Distribution and Approval Distribution to limit the amount of peers we send messages to and handle view updates.

Peer Set Manager

TODO

Protocol

Functionality

Jobs, if any

Runtime API

The Runtime API subsystem is responsible for providing a single point of access to runtime state data via a set of pre-determined queries. This prevents shared ownership of a blockchain client resource by providing

Protocol

Input: RuntimeApiMessage

Output: None

Functionality

On receipt of RuntimeApiMessage::Request(relay_parent, request), answer the request using the post-state of the relay_parent provided and provide the response to the side-channel embedded within the request.

Jobs

TODO Don't limit requests based on parent hash, but limit caching. No caching should be done for any requests on relay_parents that are not active based on ActiveLeavesUpdate messages. Maybe with some leeway for things that have just been stopped.

Chain API

The Chain API subsystem is responsible for providing a single point of access to chain state data via a set of pre-determined queries.

Protocol

Input: ChainApiMessage

Output: None

Functionality

On receipt of ChainApiMessage, answer the request and provide the response to the side-channel embedded within the request.

Currently, the following requests are supported:

  • Block hash to number
  • Block hash to header
  • Block weight
  • Finalized block number to hash
  • Last finalized block number
  • Ancestors

Chain Selection Subsystem

This subsystem implements the necessary metadata for the implementation of the chain selection portion of the protocol.

The subsystem wraps a database component which maintains a view of the unfinalized chain and records the properties of each block: whether the block is viable, whether it is stagnant, and whether it is reverted. It should also maintain an updated set of active leaves in accordance with this view, which should be cheap to query. Leaves are ordered descending first by weight and then by block number.

This subsystem needs to update its information on the unfinalized chain:

  • On every leaf-activated signal
  • On every block-finalized signal
  • On every ChainSelectionMessage::Approve
  • On every ChainSelectionMessage::RevertBlocks
  • Periodically, to detect stagnation.

Simple implementations of these updates do O(n_unfinalized_blocks) disk operations. If the amount of unfinalized blocks is relatively small, the updates should not take very much time. However, in cases where there are hundreds or thousands of unfinalized blocks the naive implementations of these update algorithms would have to be replaced with more sophisticated versions.

OverseerSignal::ActiveLeavesUpdate

Determine all new blocks implicitly referenced by any new active leaves and add them to the view. Update the set of viable leaves accordingly. The weights of imported blocks can be determined by the ChainApiMessage::BlockWeight.

OverseerSignal::BlockFinalized

Delete data for all orphaned chains and update all metadata descending from the new finalized block accordingly, along with the set of viable leaves. Note that finalizing a reverted or stagnant block means that the descendants of those blocks may lose that status because the definitions of those properties don't include the finalized chain. Update the set of viable leaves accordingly.

ChainSelectionMessage::Approved

Update the approval status of the referenced block. If the block was stagnant and thus non-viable and is now viable, then the metadata of all of its descendants needs to be updated as well, as they may no longer be stagnant either. Update the set of viable leaves accordingly.

ChainSelectionMessage::Leaves

Gets all leaves of the chain, i.e. block hashes that are suitable to build upon and have no suitable children. Supplies the leaves in descending order by score.

ChainSelectionMessage::BestLeafContaining

If the required block is unknown or not viable, then return None. Iterate over all leaves in order of descending weight, returning the first leaf containing the required block in its chain, and None otherwise.

ChainSelectionMessage::RevertBlocks

This message indicates that a dispute has concluded against a parachain block candidate. The message passes along a vector containing the block number and block hash of each block where the disputed candidate was included. The passed blocks will be marked as reverted, and their descendants will be marked as non-viable.

Periodically

Detect stagnant blocks and apply the stagnant definition to all descendants. Update the set of viable leaves accordingly.

PVF Pre-checker

The PVF pre-checker is a subsystem that is responsible for watching the relay chain for new PVFs that require pre-checking. Head over to overview for the PVF pre-checking process overview.

Protocol

There is no dedicated input mechanism for PVF pre-checker. Instead, PVF pre-checker looks on the ActiveLeavesUpdate event stream for work.

This subsytem does not produce any output messages either. The subsystem will, however, send messages to the Runtime API subsystem to query for the pending PVFs and to submit votes. In addition to that, it will also communicate with Candidate Validation Subsystem to request PVF pre-check.

Functionality

If the node is running in a collator mode, this subsystem will be disabled. The PVF pre-checker subsystem keeps track of the PVFs that are relevant for the subsystem.

To be relevant for the subsystem, a PVF must be returned by the pvfs_require_precheck runtime API in any of the active leaves. If the PVF is not present in any of the active leaves, it ceases to be relevant.

When a PVF just becomes relevant, the subsystem will send a message to the Candidate Validation subsystem asking for the pre-check.

Upon receving a message from the candidate-validation subsystem, the pre-checker will note down that the PVF has its judgement and will also sign and submit a PvfCheckStatement via the submit_pvf_check_statement runtime API. In case, a judgement was received for a PVF that is no longer in view it is ignored.

Since a vote only is valid during one session, the subsystem will have to resign and submit the statements for the new session. The new session is assumed to be started if at least one of the leaves has a greater session index that was previously observed in any of the leaves.

The subsystem tracks all the statements that it submitted within a session. If for some reason a PVF became irrelevant and then becomes relevant again, the subsystem will not submit a new statement for that PVF within the same session.

If the node is not in the active validator set, it will still perform all the checks. However, it will only submit the check statements when the node is in the active validator set.

Rejecting failed PVFs

It is possible that the candidate validation was not able to check the PVF, e.g. if it timed out. In that case, the PVF pre-checker will vote against it. This is considered safe, as there is no slashing for being on the wrong side of a pre-check vote.

Rejecting instead of abstaining is better in several ways:

  1. Conclusion is reached faster - we have actual votes, instead of relying on a timeout.
  2. Being strict in pre-checking makes it safer to be more lenient in preparation errors afterwards. Hence we have more leeway in avoiding raising dubious disputes, without making things less secure.

Also, if we only abstain, an attacker can specially craft a PVF wasm blob so that it will fail on e.g. 50% of the validators. In that case a supermajority will never be reached and the vote will repeat multiple times, most likely with the same result (since all votes are cleared on a session change). This is avoided by rejecting failed PVFs, and by only requiring 1/3 of validators to reject a PVF to reach a decision.

Note on Disputes

Having a pre-checking phase allows us to make certain assumptions later when preparing the PVF for execution. If a runtime passed pre-checking, then we know that the runtime should be valid, and therefore any issue during preparation for execution can be assumed to be a local problem on the current node.

For this reason, even deterministic preparation errors should not trigger disputes. And since we do not dispute as a result of the pre-checking phase, as stated above, it should be impossible for preparation in general to result in disputes.

Type Definitions

This section of the guide provides type definitions of various categories.

V1 Overview

Diagrams are rendered in high resolution; open them in a separate tab to see full scale.

These data types are defined in polkadot/primitives/src/v1.rs:

%3CandidateDescriptorCandidateDescriptor<H = Hash>para_idIdrelay_parentHcollatorCollatorIdpersisted_validation_data_hashHashpov_hashHasherasure_rootHashsignatureCollatorSignatureIdpolkadot_parachain::primitives::IdCandidateDescriptor:para_id->Id:wPoVHashHashCandidateDescriptor:pov_hash->PoVHashCollatorIdpolkadot_primitives::v2::CollatorIdCandidateDescriptor:collator->CollatorId:wPersistedValidationDataHashHashCandidateDescriptor:persisted_validation_data_hash->PersistedValidationDataHashPoVPoVblock_dataBlockDataPoVHash->PoV:namePersistedValidationDataPersistedValidationData<N = BlockNumber>parent_headHeadDatablock_numberNrelay_parent_storage_rootHashmax_pov_sizeu32PersistedValidationDataHash->PersistedValidationData:namepolkadot_parachain::primitives::BlockDatapolkadot_parachain::primitives::BlockDataPoV:block_data->polkadot_parachain::primitives::BlockData:wCandidateReceiptCandidateReceipt<H = Hash>descriptorCandidateDescriptor<H>commitments_hashHashCandidateReceipt:descriptor->CandidateDescriptor:nameCandidateCommitmentsHashHashCandidateReceipt:commitments_hash->CandidateCommitmentsHashCandidateCommitmentsCandidateCommitments<N = BlockNumber>upward_messagesVec<UpwardMessage>horizontal_messagesVec<OutboundHrmpMessage<Id>>new_validation_codeOption<ValidationCode>head_dataHeadDataprocessed_downward_messagesu32hrmp_watermarkNCandidateCommitmentsHash->CandidateCommitments:nameCandidateHashHashCandidateHash->CandidateReceipt:nameHeadDatapolkadot_parachain::primitives::HeadDataCandidateCommitments:head_data->HeadData:wpolkadot_parachain::primitives::UpwardMessagepolkadot_parachain::primitives::UpwardMessageCandidateCommitments:upward_messages->polkadot_parachain::primitives::UpwardMessage:wpolkadot_core_primitives::v2::OutboundHrmpMessagepolkadot_core_primitives::v2::OutboundHrmpMessageCandidateCommitments:horizontal_messages->polkadot_core_primitives::v2::OutboundHrmpMessage:wpolkadot_parachain::primitives::Idpolkadot_parachain::primitives::IdCandidateCommitments:horizontal_messages->polkadot_parachain::primitives::Id:wpolkadot_parachain::primitives::ValidationCodepolkadot_parachain::primitives::ValidationCodeCandidateCommitments:new_validation_code->polkadot_parachain::primitives::ValidationCode:wFullCandidateReceiptFullCandidateReceipt<H = Hash, N = BlockNumber>innerCandidateReceipt<H>validation_dataValidationData<N>FullCandidateReceipt:inner->CandidateReceipt:nameValidationDataValidationData<N = BlockNumber>persistedPersistedValidationData<N>transientTransientValidationData<N>FullCandidateReceipt:validation_data->ValidationData:nameValidationData:persisted->PersistedValidationData:nameTransientValidationDataTransientValidationData<N = BlockNumber>max_code_sizeu32max_head_data_sizeu32balanceBalancecode_upgrade_allowedOption<N>dmq_lengthu32ValidationData:transient->TransientValidationData:nameCommittedCandidateReceiptCommittedCandidateReceipt<H = Hash>descriptorCandidateDescriptor<H>commitmentsCandidateCommitmentsCommittedCandidateReceipt:descriptor->CandidateDescriptor:nameCommittedCandidateReceipt:commitments->CandidateCommitments:namePersistedValidationData:parent_head->HeadData:wpolkadot_core_primitives::v2::Balancepolkadot_core_primitives::v2::BalanceTransientValidationData:balance->polkadot_core_primitives::v2::Balance:wBackedCandidateBackedCandidate<H = Hash>candidateCommittedCandidateReceipt<H>validity_votesVec<ValidityAttestation>validator_indicesBitVecBackedCandidate:candidate->CommittedCandidateReceipt:namepolkadot_primitives:v0:ValidityAttestationpolkadot_primitives:v0:ValidityAttestationBackedCandidate:validity_votes->polkadot_primitives:v0:ValidityAttestation:wCoreIndexCoreIndex0u32GroupIndexGroupIndex0u32ParathreadClaimParathreadClaim0Id1CollatorIdParathreadClaim:0->Id:wParathreadClaim:1->CollatorId:wMessageQueueChainLink(prev_head, B, H(M))See doc of AbridgedHrmpChannel::mqc_headMQCHashHashMQCHash->MessageQueueChainLinkParathreadEntryParathreadEntryclaimParathreadClaimretriesu32ParathreadEntry:claim->ParathreadClaim:nameCoreOccupiedenum CoreOccupiedParathread(ParathreadEntry)ParachainCoreOccupied:parathread->ParathreadEntry:nameAvailableDataAvailableDatapovArc<PoV>validation_dataPersistedValidationDataAvailableData:pov->PoV:nameAvailableData:validation_data->PersistedValidationData:nameGroupRotationInfoGroupRotationInfo<N = BlockNumber>session_start_blockNgroup_rotation_frequencyNnowNOccupiedCoreOccupiedCore<H = Hash, N = BlockNumber>next_up_on_availableOption<ScheduledCore>occupied_sinceNtime_out_atNnext_up_on_time_outOption<ScheduledCore>availabilityBitVecgroup_responsibleGroupIndexcandidate_hashCandidateHashcandidate_descriptorCandidateDescriptorOccupiedCore:candidate_descriptor->CandidateDescriptor:nameOccupiedCore:candidate_hash->CandidateHashOccupiedCore:group_responsible->GroupIndexScheduledCoreScheduledCorepara_idIdcollatorOption<CollatorId>OccupiedCore:next_up_on_available->ScheduledCore:nameOccupiedCore:next_up_on_time_out->ScheduledCore:nameScheduledCore:para_id->Id:wScheduledCore:collator->CollatorId:wCoreStateenum CoreState<H = Hash, N = BlockNumber>Occupied(OccupiedCore<H, N>)Scheduled(ScheduledCore)FreeCoreState:occupied->OccupiedCore:nameCoreState:scheduled->ScheduledCore:nameCandidateEventenum CandidateEvent<H = Hash>CandidateBacked(CandidateReceipt<H>, HeadData)CandidateIncluded(CandidateReceipt<H>, HeadData)CandidateTimedOut(CandidateReceipt<H>, HeadData)CandidateEvent:e->CandidateReceipt:nameCandidateEvent:e->HeadData:wSessionInfoSessionInfovalidatorsVec<ValidatorId>discovery_keysVec<AuthorityDiscoveryId>assignment_keysVec<AssignmentId>validator_groupsVec<Vec<ValidatorIndex>>n_coresu32zeroth_delay_tranche_widthu32relay_vrf_modulo_samplesu32n_delay_tranchesu32no_show_slotsu32needed_approvalsu32ValidatorIdpolkadot_primitives::v2::ValidatorIdSessionInfo:validators->ValidatorId:wAuthorityDiscoveryIdsp_authority_discovery::AuthorityIdSessionInfo:discovery_keys->AuthorityDiscoveryId:wValidatorIndexpolkadot_primitives::v2::ValidatorIndexSessionInfo:validator_groups->ValidatorIndex:wAbridgedHostConfigurationAbridgedHostConfigurationmax_code_sizeu32max_head_data_sizeu32max_upward_queue_countu32max_upward_queue_sizeu32max_upward_message_sizeu32max_upward_messages_num_per_candidateu32hrmp_max_message_num_per_candidateu32validation_upgrade_cooldownBlockNumbervalidation_upgrade_delayBlockNumberAbridgedHrmpChannelAbridgedHrmpChannelmax_capacityu32max_total_sizeu32max_message_sizeu32msg_countu32total_sizeu32mqc_headOption<Hash>AbridgedHrmpChannel:mqc_head->MQCHash

These data types are defined in polkadot/parachain/src/primitives.rs:

%3HeadDataHeadData0Vec<u8>ValidationCodeValidationCode0Vec<u8>BlockDataBlockData0Vec<u8>IdId0u32SiblingSibling0IdSibling:0->Id:nameHrmpChannelIdHrmpChannelIdsenderIdrecipientIdHrmpChannelId:e->Id:nameValidationParamsValidationParamsparent_headHeadDatablock_dataBlockDatarelay_parent_numberRelayChainBlockNumberrelay_parent_storage_rootHashValidationParams:parent_head->HeadData:nameValidationParams:block_data->BlockData:nameRelayChainBlockNumberpolkadot_core_primitives::BlockNumberValidationParams:relay_parent_number->RelayChainBlockNumber:wValidationResultValidationResulthead_dataHeadDatanew_validation_codeOption<ValidationCode>upward_messagesVec<UpwardMessage>horizontal_messagesVec<OutboundHrmpMessage<Id>>processed_downward_messagesu32hrmp_watermarkRelayChainBlockNumberValidationResult:head_data->HeadData:nameValidationResult:new_validation_code->ValidationCode:nameValidationResult:horizontal_messages->Id:nameValidationResult:hrmp_watermark->RelayChainBlockNumber:wUpwardMessageVec<u8>ValidationResult:upward_messages->UpwardMessage:wOutboundHrmpMessagepolkadot_core_primitives::OutboundHrmpMessageValidationResult:horizontal_messages->OutboundHrmpMessage:w

Candidate Types

Para candidates are some of the most common types, both within the runtime and on the Node-side. Candidates are the fundamental datatype for advancing parachains, encapsulating the collator's signature, the context of the parablock, the commitments to the output, and a commitment to the data which proves it valid.

In a way, this entire guide is about these candidates: how they are scheduled, constructed, backed, included, and challenged.

This section will describe the base candidate type, its components, and variants that contain extra data.

Para Id

A unique 32-bit identifier referring to a specific para (chain or thread). The relay-chain runtime guarantees that ParaIds are unique for the duration of any session, but recycling and reuse over a longer period of time is permitted.

#![allow(unused)]
fn main() {
struct ParaId(u32);
}

Candidate Receipt

Much info in a FullCandidateReceipt is duplicated from the relay-chain state. When the corresponding relay-chain state is considered widely available, the Candidate Receipt should be favored over the FullCandidateReceipt.

Examples of situations where the state is readily available includes within the scope of work done by subsystems working on a given relay-parent, or within the logic of the runtime importing a backed candidate.

#![allow(unused)]
fn main() {
/// A candidate-receipt.
struct CandidateReceipt {
	/// The descriptor of the candidate.
	descriptor: CandidateDescriptor,
	/// The hash of the encoded commitments made as a result of candidate execution.
	commitments_hash: Hash,
}
}

Full Candidate Receipt

This is the full receipt type. The PersistedValidationData are technically redundant with the inner.relay_parent, which uniquely describes the block in the blockchain from whose state these values are derived. The CandidateReceipt variant is often used instead for this reason.

However, the Full Candidate Receipt type is useful as a means of avoiding the implicit dependency on availability of old blockchain state. In situations such as availability and approval, having the full description of the candidate within a self-contained struct is convenient.

#![allow(unused)]
fn main() {
/// All data pertaining to the execution of a para candidate.
struct FullCandidateReceipt {
	inner: CandidateReceipt,
	validation_data: PeristedValidationData,
}
}

Committed Candidate Receipt

This is a variant of the candidate receipt which includes the commitments of the candidate receipt alongside the descriptor. This should be favored over the Candidate Receipt in situations where the candidate is not going to be executed but the actual data committed to is important. This is often the case in the backing phase.

The hash of the committed candidate receipt will be the same as the corresponding Candidate Receipt, because it is computed by first hashing the encoding of the commitments to form a plain Candidate Receipt.

#![allow(unused)]
fn main() {
/// A candidate-receipt with commitments directly included.
struct CommittedCandidateReceipt {
	/// The descriptor of the candidate.
	descriptor: CandidateDescriptor,
	/// The commitments of the candidate receipt.
	commitments: CandidateCommitments,
}
}

Candidate Descriptor

This struct is pure description of the candidate, in a lightweight format.

#![allow(unused)]
fn main() {
/// A unique descriptor of the candidate receipt.
struct CandidateDescriptor {
	/// The ID of the para this is a candidate for.
	para_id: ParaId,
	/// The hash of the relay-chain block this is executed in the context of.
	relay_parent: Hash,
	/// The collator's sr25519 public key.
	collator: CollatorId,
	/// The blake2-256 hash of the persisted validation data. These are extra parameters
	/// derived from relay-chain state that influence the validity of the block which
	/// must also be kept available for approval checkers.
	persisted_validation_data_hash: Hash,
	/// The blake2-256 hash of the `pov-block`.
	pov_hash: Hash,
	/// The root of a block's erasure encoding Merkle tree.
	erasure_root: Hash,
	/// Signature on blake2-256 of components of this receipt:
	/// The parachain index, the relay parent, the validation data hash, and the `pov_hash`.
	signature: CollatorSignature,
	/// Hash of the para header that is being generated by this candidate.
	para_head: Hash,
	/// The blake2-256 hash of the validation code bytes.
	validation_code_hash: ValidationCodeHash,
}
}

ValidationParams

#![allow(unused)]
fn main() {
/// Validation parameters for evaluating the parachain validity function.
pub struct ValidationParams {
	/// Previous head-data.
	pub parent_head: HeadData,
	/// The collation body.
	pub block_data: BlockData,
	/// The current relay-chain block number.
	pub relay_parent_number: RelayChainBlockNumber,
	/// The relay-chain block's storage root.
	pub relay_parent_storage_root: Hash,
}
}

PersistedValidationData

The validation data provides information about how to create the inputs for validation of a candidate. This information is derived from the chain state and will vary from para to para, although some of the fields may be the same for every para.

Since this data is used to form inputs to the validation function, it needs to be persisted by the availability system to avoid dependence on availability of the relay-chain state.

Furthermore, the validation data acts as a way to authorize the additional data the collator needs to pass to the validation function. For example, the validation function can check whether the incoming messages (e.g. downward messages) were actually sent by using the data provided in the validation data using so called MQC heads.

Since the commitments of the validation function are checked by the relay-chain, approval checkers can rely on the invariant that the relay-chain only includes para-blocks for which these checks have already been done. As such, there is no need for the validation data used to inform validators and collators about the checks the relay-chain will perform to be persisted by the availability system.

The PersistedValidationData should be relatively lightweight primarily because it is constructed during inclusion for each candidate and therefore lies on the critical path of inclusion.

#![allow(unused)]
fn main() {
struct PersistedValidationData {
	/// The parent head-data.
	parent_head: HeadData,
	/// The relay-chain block number this is in the context of. This informs the collator.
	relay_parent_number: BlockNumber,
	/// The relay-chain block storage root this is in the context of.
	relay_parent_storage_root: Hash,
	/// The list of MQC heads for the inbound channels paired with the sender para ids. This
	/// vector is sorted ascending by the para id and doesn't contain multiple entries with the same
	/// sender.
	///
	/// The HRMP MQC heads will be used by the validation function to authorize the input messages passed
	/// by the collator.
	hrmp_mqc_heads: Vec<(ParaId, Hash)>,
	/// The maximum legal size of a POV block, in bytes.
	pub max_pov_size: u32,
}
}

HeadData

Head data is a type-safe abstraction around bytes (Vec<u8>) for the purposes of representing heads of parachains.

#![allow(unused)]
fn main() {
struct HeadData(Vec<u8>);
}

Candidate Commitments

The execution and validation of parachain candidates produces a number of values which either must be committed to blocks on the relay chain or committed to the state of the relay chain.

#![allow(unused)]
fn main() {
/// Commitments made in a `CandidateReceipt`. Many of these are outputs of validation.
#[derive(PartialEq, Eq, Clone, Encode, Decode)]
#[cfg_attr(feature = "std", derive(Debug, Default))]
struct CandidateCommitments {
	/// Messages directed to other paras routed via the relay chain.
	horizontal_messages: Vec<OutboundHrmpMessage>,
	/// Messages destined to be interpreted by the Relay chain itself.
	upward_messages: Vec<UpwardMessage>,
	/// New validation code.
	new_validation_code: Option<ValidationCode>,
	/// The head-data produced as a result of execution.
	head_data: HeadData,
	/// The number of messages processed from the DMQ.
	processed_downward_messages: u32,
	/// The mark which specifies the block number up to which all inbound HRMP messages are processed.
	hrmp_watermark: BlockNumber,
}
}

Signing Context

This struct provides context to signatures by combining with various payloads to localize the signature to a particular session index and relay-chain hash. Having these fields included in the signature makes misbehavior attribution much simpler.

#![allow(unused)]
fn main() {
struct SigningContext {
	/// The relay-chain block hash this signature is in the context of.
	parent_hash: Hash,
	/// The session index this signature is in the context of.
	session_index: SessionIndex,
}
}

Backing Types

Candidates go through many phases before being considered included in a fork of the relay chain and eventually accepted.

These types describe the data used in the backing phase. Some are sent over the wire within subsystems, and some are simply included in the relay-chain block.

Validity Attestation

An attestation of validity for a candidate, used as part of a backing. Both the Seconded and Valid statements are considered attestations of validity. This structure is only useful where the candidate referenced is apparent.

#![allow(unused)]
fn main() {
enum ValidityAttestation {
  /// Implicit validity attestation by issuing.
  /// This corresponds to issuance of a `Seconded` statement.
  Implicit(ValidatorSignature),
  /// An explicit attestation. This corresponds to issuance of a
  /// `Valid` statement.
  Explicit(ValidatorSignature),
}
}

Signed Wrapper

There are a few distinct types which we desire to sign, and validate the signatures of. Instead of duplicating this work, we extract a signed wrapper.

/// A signed type which encapsulates the common desire to sign some data and validate a signature.
///
/// Note that the internal fields are not public; they are all accessable by immutable getters.
/// This reduces the chance that they are accidentally mutated, invalidating the signature.
struct Signed<Payload, RealPayload=Payload> {
    /// The payload is part of the signed data. The rest is the signing context,
    /// which is known both at signing and at validation.
    payload: Payload,
    /// The index of the validator signing this statement.
    validator_index: ValidatorIndex,
    /// The signature by the validator of the signed payload.
    signature: ValidatorSignature,
}

impl<Payload: EncodeAs<RealPayload>, RealPayload: Encode> Signed<Payload, RealPayload> {
    fn sign(payload: Payload, context: SigningContext, index: ValidatorIndex, key: ValidatorPair) -> Signed<Payload, RealPayload> { ... }
    fn validate(&self, context: SigningContext, key: ValidatorId) -> bool { ... }
}

Note the presence of the SigningContext in the signatures of the sign and validate methods. To ensure cryptographic security, the actual signed payload is always the SCALE encoding of (payload.into(), signing_context). Including the signing context prevents replay attacks.

EncodeAs is a helper trait with a blanket impl which ensures that any T can EncodeAs<T>. Therefore, for the generic case where RealPayload = Payload, it changes nothing. However, we impl EncodeAs<CompactStatement> for Statement, which helps efficiency.

Statement Type

The Candidate Backing subsystem issues and signs these after candidate validation.

#![allow(unused)]
fn main() {
/// A statement about the validity of a parachain candidate.
enum Statement {
  /// A statement about a new candidate being seconded by a validator. This is an implicit validity vote.
  ///
  /// The main semantic difference between `Seconded` and `Valid` comes from the fact that every validator may
  /// second only 1 candidate; this places an upper bound on the total number of candidates whose validity
  /// needs to be checked. A validator who seconds more than 1 parachain candidate per relay head is subject
  /// to slashing.
  Seconded(CommittedCandidateReceipt),
  /// A statement about the validity of a candidate, based on candidate's hash.
  Valid(Hash),
}

/// A statement about the validity of a parachain candidate.
///
/// This variant should only be used in the production of `SignedStatement`s. The only difference between
/// this enum and `Statement` is that the `Seconded` variant contains a `Hash` instead of a `CandidateReceipt`.
/// The rationale behind the difference is that the signature should always be on the hash instead of the
/// full data, as this lowers the requirement for checking while retaining necessary cryptographic properties
enum CompactStatement {
  /// A statement about a new candidate being seconded by a validator. This is an implicit validity vote.
  Seconded(Hash),
  /// A statement about the validity of a candidate, based on candidate's hash.
  Valid(Hash),
}
}

CompactStatement exists because a CandidateReceipt includes HeadData, which does not have a bounded size.

Signed Statement Type

A statement which has been cryptographically signed by a validator.

#![allow(unused)]
fn main() {
/// A signed statement, containing the committed candidate receipt in the `Seconded` variant.
pub type SignedFullStatement = Signed<Statement, CompactStatement>;

/// A signed statement, containing only the hash.
pub type SignedStatement = Signed<CompactStatement>;
}

Munging the signed Statement into a CompactStatement before signing allows the candidate receipt itself to be omitted when checking a signature on a Seconded statement.

Backed Candidate

An CommittedCandidateReceipt along with all data necessary to prove its backing. This is submitted to the relay-chain to process and move along the candidate to the pending-availability stage.

#![allow(unused)]
fn main() {
struct BackedCandidate {
  candidate: CommittedCandidateReceipt,
  validity_votes: Vec<ValidityAttestation>,
  // the indices of validators who signed the candidate within the group. There is no need to include
  // bit for any validators who are not in the group, so this is more compact.
  // The number of bits is the number of validators in the group.
  //
  // the group should be apparent from context.
  validator_indices: BitVec,
}

struct BackedCandidates(Vec<BackedCandidate>); // sorted by para-id.
}

Availability

One of the key roles of validators is to ensure availability of all data necessary to validate candidates for the duration of a challenge period. This is done via an erasure-coding of the data to keep available.

Signed Availability Bitfield

A bitfield signed by a particular validator about the availability of pending candidates.

#![allow(unused)]
fn main() {
type SignedAvailabilityBitfield = Signed<Bitvec>;

struct Bitfields(Vec<(SignedAvailabilityBitfield)>), // bitfields sorted by validator index, ascending
}

Semantics

A SignedAvailabilityBitfield represents the view from a particular validator's perspective. Each bit in the bitfield corresponds to a single availability core. A 1 bit indicates that the validator believes the following statements to be true for a core:

  • the availability core is occupied
  • there exists a CommittedCandidateReceipt corresponding to that core. In other words, that para has a block in progress.
  • the validator's Availability Store contains a chunk of that parablock's PoV.

In other words, it is the transpose of OccupiedCore::availability.

Proof-of-Validity

Often referred to as PoV, this is a type-safe wrapper around bytes (Vec<u8>) when referring to data that acts as a stateless-client proof of validity of a candidate, when used as input to the validation function of the para.

#![allow(unused)]
fn main() {
struct PoV(Vec<u8>);
}

Available Data

This is the data we want to keep available for each candidate included in the relay chain. This is the PoV of the block, as well as the PersistedValidationData

#![allow(unused)]
fn main() {
struct AvailableData {
    /// The Proof-of-Validation of the candidate.
    pov: Arc<PoV>,
    /// The persisted validation data used to check the candidate.
    validation_data: PersistedValidationData,
}
}

TODO: With XCMP, we also need to keep available the outgoing messages as a result of para-validation.

Erasure Chunk

The AvailableData is split up into an erasure-coding as part of the availability process. Each validator gets a chunk. This describes one of those chunks, along with its proof against a merkle root hash, which should be apparent from context, and is the erasure_root field of a CandidateDescriptor.

#![allow(unused)]
fn main() {
struct ErasureChunk {
    /// The erasure-encoded chunk of data belonging to the candidate block.
    chunk: Vec<u8>,
    /// The index of this erasure-encoded chunk of data.
    index: u32,
    /// Proof for this chunk's branch in the Merkle tree.
    proof: Vec<Vec<u8>>,
}
}

Overseer Protocol

This chapter contains message types sent to and from the overseer, and the underlying subsystem message types that are transmitted using these.

Overseer Signal

Signals from the overseer to a subsystem to request change in execution that has to be obeyed by the subsystem.

#![allow(unused)]
fn main() {
enum OverseerSignal {
  /// Signal about a change in active leaves.
  ActiveLeavesUpdate(ActiveLeavesUpdate),
  /// Signal about a new best finalized block.
  BlockFinalized(Hash),
  /// Conclude all operation.
  Conclude,
}
}

All subsystems have their own message types; all of them need to be able to listen for overseer signals as well. There are currently two proposals for how to handle that with unified communication channels:

  1. Retaining the OverseerSignal definition above, add enum FromOrchestra<T> {Signal(OverseerSignal), Message(T)}.
  2. Add a generic varint to OverseerSignal: Message(T).

Either way, there will be some top-level type encapsulating messages from the overseer to each subsystem.

Active Leaves Update

Indicates a change in active leaves. Activated leaves should have jobs, whereas deactivated leaves should lead to winding-down of work based on those leaves.

#![allow(unused)]
fn main() {
enum LeafStatus {
    // A leaf is fresh when it's the first time the leaf has been encountered.
    // Most leaves should be fresh.
    Fresh,
    // A leaf is stale when it's encountered for a subsequent time. This will
    // happen when the chain is reverted or the fork-choice rule abandons some
    // chain.
    Stale,
}

struct ActiveLeavesUpdate {
    activated: [(Hash, Number, LeafStatus)], // in practice, these should probably be a SmallVec
    deactivated: [Hash],
}
}

All Messages

A message type tying together all message types that are used across Subsystems.

#![allow(unused)]
fn main() {
enum AllMessages {
    CandidateValidation(CandidateValidationMessage),
    CandidateBacking(CandidateBackingMessage),
    ChainApi(ChainApiMessage),
    CollatorProtocol(CollatorProtocolMessage),
    StatementDistribution(StatementDistributionMessage),
    AvailabilityDistribution(AvailabilityDistributionMessage),
    AvailabilityRecovery(AvailabilityRecoveryMessage),
    BitfieldDistribution(BitfieldDistributionMessage),
    BitfieldSigning(BitfieldSigningMessage),
    Provisioner(ProvisionerMessage),
    RuntimeApi(RuntimeApiMessage),
    AvailabilityStore(AvailabilityStoreMessage),
    NetworkBridge(NetworkBridgeMessage),
    CollationGeneration(CollationGenerationMessage),
    ApprovalVoting(ApprovalVotingMessage),
    ApprovalDistribution(ApprovalDistributionMessage),
    GossipSupport(GossipSupportMessage),
    DisputeCoordinator(DisputeCoordinatorMessage),
    ChainSelection(ChainSelectionMessage),
    PvfChecker(PvfCheckerMessage),
}
}

Approval Voting Message

Messages received by the approval voting subsystem.

#![allow(unused)]
fn main() {
enum AssignmentCheckResult {
    // The vote was accepted and should be propagated onwards.
    Accepted,
    // The vote was valid but duplicate and should not be propagated onwards.
    AcceptedDuplicate,
    // The vote was valid but too far in the future to accept right now.
    TooFarInFuture,
    // The vote was bad and should be ignored, reporting the peer who propagated it.
    Bad(AssignmentCheckError),
}

pub enum AssignmentCheckError {
    UnknownBlock(Hash),
    UnknownSessionIndex(SessionIndex),
    InvalidCandidateIndex(CandidateIndex),
    InvalidCandidate(CandidateIndex, CandidateHash),
    InvalidCert(ValidatorIndex),
    Internal(Hash, CandidateHash),
}

enum ApprovalCheckResult {
    // The vote was accepted and should be propagated onwards.
    Accepted,
    // The vote was bad and should be ignored, reporting the peer who propagated it.
    Bad(ApprovalCheckError),
}

pub enum ApprovalCheckError {
    UnknownBlock(Hash),
    UnknownSessionIndex(SessionIndex),
    InvalidCandidateIndex(CandidateIndex),
    InvalidValidatorIndex(ValidatorIndex),
    InvalidCandidate(CandidateIndex, CandidateHash),
    InvalidSignature(ValidatorIndex),
    NoAssignment(ValidatorIndex),
    Internal(Hash, CandidateHash),
}

enum ApprovalVotingMessage {
    /// Check if the assignment is valid and can be accepted by our view of the protocol.
    /// Should not be sent unless the block hash is known.
    CheckAndImportAssignment(
        IndirectAssignmentCert,
        CandidateIndex, // The index of the candidate included in the block.
        ResponseChannel<AssignmentCheckResult>,
    ),
    /// Check if the approval vote is valid and can be accepted by our view of the
    /// protocol.
    ///
    /// Should not be sent unless the block hash within the indirect vote is known.
    CheckAndImportApproval(
        IndirectSignedApprovalVote,
        ResponseChannel<ApprovalCheckResult>,
    ),
    /// Returns the highest possible ancestor hash of the provided block hash which is
    /// acceptable to vote on finality for. Along with that, return the lists of candidate hashes
    /// which appear in every block from the (non-inclusive) base number up to (inclusive) the specified
    /// approved ancestor.
    /// This list starts from the highest block (the approved ancestor itself) and moves backwards
    /// towards the base number.
    ///
    /// The base number is typically the number of the last finalized block, but in GRANDPA it is
    /// possible for the base to be slightly higher than the last finalized block.
    ///
    /// The `BlockNumber` provided is the number of the block's ancestor which is the
    /// earliest possible vote.
    ///
    /// It can also return the same block hash, if that is acceptable to vote upon.
    /// Return `None` if the input hash is unrecognized.
    ApprovedAncestor {
        target_hash: Hash,
        base_number: BlockNumber,
        rx: ResponseChannel<Option<(Hash, BlockNumber, Vec<(Hash, Vec<CandidateHash>)>)>>
    },
}
}

Approval Distribution Message

Messages received by the approval distribution subsystem.

#![allow(unused)]
fn main() {
/// Metadata about a block which is now live in the approval protocol.
struct BlockApprovalMeta {
    /// The hash of the block.
    hash: Hash,
    /// The number of the block.
    number: BlockNumber,
    /// The candidates included by the block. Note that these are not the same as the candidates that appear within the
    /// block body.
    parent_hash: Hash,
    /// The candidates included by the block. Note that these are not the same as the candidates that appear within the
    /// block body.
    candidates: Vec<CandidateHash>,
    /// The consensus slot of the block.
    slot: Slot,
    /// The session of the block.
    session: SessionIndex,
}

enum ApprovalDistributionMessage {
    /// Notify the `ApprovalDistribution` subsystem about new blocks and the candidates contained within
    /// them.
    NewBlocks(Vec<BlockApprovalMeta>),
    /// Distribute an assignment cert from the local validator. The cert is assumed
    /// to be valid, relevant, and for the given relay-parent and validator index.
    ///
    /// The `u32` param is the candidate index in the fully-included list.
    DistributeAssignment(IndirectAssignmentCert, u32),
    /// Distribute an approval vote for the local validator. The approval vote is assumed to be
    /// valid, relevant, and the corresponding approval already issued. If not, the subsystem is free to drop
    /// the message.
    DistributeApproval(IndirectSignedApprovalVote),
    /// An update from the network bridge.
    NetworkBridgeUpdate(NetworkBridgeEvent<ApprovalDistributionV1Message>),
}
}

Availability Distribution Message

Messages received by the availability distribution subsystem.

This is a network protocol that receives messages of type AvailabilityDistributionV1Message.

#![allow(unused)]
fn main() {
enum AvailabilityDistributionMessage {
      /// Incoming network request for an availability chunk.
      ChunkFetchingRequest(IncomingRequest<req_res_v1::ChunkFetchingRequest>),
      /// Incoming network request for a seconded PoV.
      PoVFetchingRequest(IncomingRequest<req_res_v1::PoVFetchingRequest>),
      /// Instruct availability distribution to fetch a remote PoV.
      ///
      /// NOTE: The result of this fetch is not yet locally validated and could be bogus.
      FetchPoV {
          /// The relay parent giving the necessary context.
          relay_parent: Hash,
          /// Validator to fetch the PoV from.
          from_validator: ValidatorIndex,
          /// Candidate hash to fetch the PoV for.
          candidate_hash: CandidateHash,
          /// Expected hash of the PoV, a PoV not matching this hash will be rejected.
          pov_hash: Hash,
          /// Sender for getting back the result of this fetch.
          ///
          /// The sender will be canceled if the fetching failed for some reason.
          tx: oneshot::Sender<PoV>,
      },
}
}

Availability Recovery Message

Messages received by the availability recovery subsystem.

#![allow(unused)]
fn main() {
enum RecoveryError {
    Invalid,
    Unavailable,
}
enum AvailabilityRecoveryMessage {
    /// Recover available data from validators on the network.
    RecoverAvailableData(
        CandidateReceipt,
        SessionIndex,
        Option<GroupIndex>, // Backing validator group to request the data directly from.
        ResponseChannel<Result<AvailableData, RecoveryError>>,
    ),
}
}

Availability Store Message

Messages to and from the availability store.

#![allow(unused)]
fn main() {
pub enum AvailabilityStoreMessage {
	/// Query a `AvailableData` from the AV store.
	QueryAvailableData(CandidateHash, oneshot::Sender<Option<AvailableData>>),

	/// Query whether a `AvailableData` exists within the AV Store.
	///
	/// This is useful in cases when existence
	/// matters, but we don't want to necessarily pass around multiple
	/// megabytes of data to get a single bit of information.
	QueryDataAvailability(CandidateHash, oneshot::Sender<bool>),

	/// Query an `ErasureChunk` from the AV store by the candidate hash and validator index.
	QueryChunk(CandidateHash, ValidatorIndex, oneshot::Sender<Option<ErasureChunk>>),

	/// Get the size of an `ErasureChunk` from the AV store by the candidate hash.
	QueryChunkSize(CandidateHash, oneshot::Sender<Option<usize>>),

	/// Query all chunks that we have for the given candidate hash.
	QueryAllChunks(CandidateHash, oneshot::Sender<Vec<ErasureChunk>>),

	/// Query whether an `ErasureChunk` exists within the AV Store.
	///
	/// This is useful in cases like bitfield signing, when existence
	/// matters, but we don't want to necessarily pass around large
	/// quantities of data to get a single bit of information.
	QueryChunkAvailability(CandidateHash, ValidatorIndex, oneshot::Sender<bool>),

	/// Store an `ErasureChunk` in the AV store.
	///
	/// Return `Ok(())` if the store operation succeeded, `Err(())` if it failed.
	StoreChunk {
		/// A hash of the candidate this chunk belongs to.
		candidate_hash: CandidateHash,
		/// The chunk itself.
		chunk: ErasureChunk,
		/// Sending side of the channel to send result to.
		tx: oneshot::Sender<Result<(), ()>>,
	},

	/// Computes and checks the erasure root of `AvailableData` before storing all of its chunks in 
	/// the AV store.
	///
	/// Return `Ok(())` if the store operation succeeded, `Err(StoreAvailableData)` if it failed.
	StoreAvailableData {
		/// A hash of the candidate this `available_data` belongs to.
		candidate_hash: CandidateHash,
		/// The number of validators in the session.
		n_validators: u32,
		/// The `AvailableData` itself.
		available_data: AvailableData,
		/// Erasure root we expect to get after chunking.
		expected_erasure_root: Hash,
		/// Sending side of the channel to send result to.
		tx: oneshot::Sender<Result<(), StoreAvailableDataError>>,
	},
}

/// The error result type of a [`AvailabilityStoreMessage::StoreAvailableData`] request.
pub enum StoreAvailableDataError {
	InvalidErasureRoot,
}
}

Bitfield Distribution Message

Messages received by the bitfield distribution subsystem. This is a network protocol that receives messages of type BitfieldDistributionV1Message.

#![allow(unused)]
fn main() {
enum BitfieldDistributionMessage {
    /// Distribute a bitfield signed by a validator to other validators.
    /// The bitfield distribution subsystem will assume this is indeed correctly signed.
    DistributeBitfield(relay_parent, SignedAvailabilityBitfield),
    /// Receive a network bridge update.
    NetworkBridgeUpdate(NetworkBridgeEvent<BitfieldDistributionV1Message>),
}
}

Bitfield Signing Message

Currently, the bitfield signing subsystem receives no specific messages.

#![allow(unused)]
fn main() {
/// Non-instantiable message type
enum BitfieldSigningMessage { }
}

Candidate Backing Message

#![allow(unused)]
fn main() {
enum CandidateBackingMessage {
  /// Requests a set of backable candidates attested by the subsystem.
  ///
  /// Each pair is (candidate_hash, candidate_relay_parent).
  GetBackedCandidates(Vec<(CandidateHash, Hash)>, oneshot::Sender<Vec<BackedCandidate>>),
  /// Note that the Candidate Backing subsystem should second the given candidate in the context of the
  /// given relay-parent (ref. by hash). This candidate must be validated using the provided PoV.
  /// The PoV is expected to match the `pov_hash` in the descriptor.
  Second(Hash, CandidateReceipt, PoV),
  /// Note a peer validator's statement about a particular candidate. Disagreements about validity must be escalated
  /// to a broader check by the Disputes Subsystem, though that escalation is deferred until the approval voting 
  /// stage to guarantee availability. Agreements are simply tallied until a quorum is reached.
  Statement(Statement),
}
}

Chain API Message

The Chain API subsystem is responsible for providing an interface to chain data.

#![allow(unused)]
fn main() {
enum ChainApiMessage {
    /// Get the block number by hash.
    /// Returns `None` if a block with the given hash is not present in the db.
    BlockNumber(Hash, ResponseChannel<Result<Option<BlockNumber>, Error>>),
    /// Request the block header by hash.
    /// Returns `None` if a block with the given hash is not present in the db.
    BlockHeader(Hash, ResponseChannel<Result<Option<BlockHeader>, Error>>),
    /// Get the cumulative weight of the given block, by hash.
    /// If the block or weight is unknown, this returns `None`.
    ///
    /// Weight is used for comparing blocks in a fork-choice rule.
    BlockWeight(Hash, ResponseChannel<Result<Option<Weight>, Error>>),
    /// Get the finalized block hash by number.
    /// Returns `None` if a block with the given number is not present in the db.
    /// Note: the caller must ensure the block is finalized.
    FinalizedBlockHash(BlockNumber, ResponseChannel<Result<Option<Hash>, Error>>),
    /// Get the last finalized block number.
    /// This request always succeeds.
    FinalizedBlockNumber(ResponseChannel<Result<BlockNumber, Error>>),
    /// Request the `k` ancestors block hashes of a block with the given hash.
    /// The response channel may return a `Vec` of size up to `k`
    /// filled with ancestors hashes with the following order:
    /// `parent`, `grandparent`, ...
    Ancestors {
        /// The hash of the block in question.
        hash: Hash,
        /// The number of ancestors to request.
        k: usize,
        /// The response channel.
        response_channel: ResponseChannel<Result<Vec<Hash>, Error>>,
    }
}
}

Chain Selection Message

Messages received by the Chain Selection subsystem

#![allow(unused)]
fn main() {
enum ChainSelectionMessage {
    /// Signal to the chain selection subsystem that a specific block has been approved.
    Approved(Hash),
    /// Request the leaves in descending order by score.
    Leaves(ResponseChannel<Vec<Hash>>),
    /// Request the best leaf containing the given block in its ancestry. Return `None` if
    /// there is no such leaf.
    BestLeafContaining(Hash, ResponseChannel<Option<Hash>>),

}
}

Collator Protocol Message

Messages received by the Collator Protocol subsystem

This is a network protocol that receives messages of type CollatorProtocolV1Message.

#![allow(unused)]
fn main() {
enum CollatorProtocolMessage {
    /// Signal to the collator protocol that it should connect to validators with the expectation
    /// of collating on the given para. This is only expected to be called once, early on, if at all,
    /// and only by the Collation Generation subsystem. As such, it will overwrite the value of
    /// the previous signal.
    ///
    /// This should be sent before any `DistributeCollation` message.
    CollateOn(ParaId),
    /// Provide a collation to distribute to validators with an optional result sender.
    ///
    /// The result sender should be informed when at least one parachain validator seconded the collation. It is also
    /// completely okay to just drop the sender.
    DistributeCollation(CandidateReceipt, PoV, Option<oneshot::Sender<CollationSecondedSignal>>),
    /// Fetch a collation under the given relay-parent for the given ParaId.
    FetchCollation(Hash, ParaId, ResponseChannel<(CandidateReceipt, PoV)>),
    /// Report a collator as having provided an invalid collation. This should lead to disconnect
    /// and blacklist of the collator.
    ReportCollator(CollatorId),
    /// Note a collator as having provided a good collation.
    NoteGoodCollation(CollatorId, SignedFullStatement),
    /// Notify a collator that its collation was seconded.
    NotifyCollationSeconded(CollatorId, Hash, SignedFullStatement),
}
}

Collation Generation Message

Messages received by the Collation Generation subsystem

This is the core interface by which collators built on top of a Polkadot node submit collations to validators. As such, these messages are not sent by any subsystem but are instead sent from outside of the overseer.

#![allow(unused)]
fn main() {
/// A function provided to the subsystem which it uses to pull new collations.
///
/// This mode of querying collations is obsoleted by `CollationGenerationMessages::SubmitCollation`
///
/// The response channel, if present, is meant to receive a `Seconded` statement as a
/// form of authentication, for collation mechanisms which rely on this for anti-spam.
type CollatorFn = Fn(Hash, PersistedValidationData) -> Future<Output = (Collation, Option<ResponseChannel<SignedStatement>>)>;

/// Configuration for the collation generator
struct CollationGenerationConfig {
    /// Collator's authentication key, so it can sign things.
    key: CollatorPair,
    /// Collation function. See [`CollatorFn`] for more details.
    collator: CollatorFn,
    /// The parachain that this collator collates for
    para_id: ParaId,
}

/// Parameters for submitting a collation
struct SubmitCollationParams {
    /// The relay-parent the collation is built against.
    relay_parent: Hash,
    /// The collation itself (PoV and commitments)
    collation: Collation,
    /// The parent block's head-data.
    parent_head: HeadData,
    /// The hash of the validation code the collation was created against.
    validation_code_hash: ValidationCodeHash,
    /// A response channel for receiving a `Seconded` message about the candidate
    /// once produced by a validator. This is not guaranteed to provide anything.
    result_sender: Option<ResponseChannel<SignedStatement>>,
}

enum CollationGenerationMessage {
    /// Initialize the collation generation subsystem
	Initialize(CollationGenerationConfig),
    /// Submit a collation to the subsystem. This will package it into a signed
    /// [`CommittedCandidateReceipt`] and distribute along the network to validators.
    ///
    /// If sent before `Initialize`, this will be ignored.
    SubmitCollation(SubmitCollationParams),
}
}

Dispute Coordinator Message

Messages received by the Dispute Coordinator subsystem

This subsystem coordinates participation in disputes, tracks live disputes, and observed statements of validators from subsystems.

#![allow(unused)]
fn main() {
enum DisputeCoordinatorMessage {
    /// Import a statement by a validator about a candidate.
    ///
    /// The subsystem will silently discard ancient statements or sets of only dispute-specific statements for
    /// candidates that are previously unknown to the subsystem. The former is simply because ancient
    /// data is not relevant and the latter is as a DoS prevention mechanism. Both backing and approval
    /// statements already undergo anti-DoS procedures in their respective subsystems, but statements
    /// cast specifically for disputes are not necessarily relevant to any candidate the system is
    /// already aware of and thus present a DoS vector. Our expectation is that nodes will notify each
    /// other of disputes over the network by providing (at least) 2 conflicting statements, of which one is either
    /// a backing or validation statement.
    ///
    /// This does not do any checking of the message signature.
    ImportStatements {
        /// The hash of the candidate.
        candidate_hash: CandidateHash,
        /// The candidate receipt itself.
        candidate_receipt: CandidateReceipt,
        /// The session the candidate appears in.
        session: SessionIndex,
        /// Triples containing the following:
        /// - A statement, either indicating validity or invalidity of the candidate.
        /// - The validator index (within the session of the candidate) of the validator casting the vote.
        /// - The signature of the validator casting the vote.
        statements: Vec<(DisputeStatement, ValidatorIndex, ValidatorSignature)>,

        /// Inform the requester once we finished importing.
        ///
        /// This is, we either discarded the votes, just record them because we
        /// casted our vote already or recovered availability for the candidate
        /// successfully.
        pending_confirmation: oneshot::Sender<ImportStatementsResult>
    },
    /// Fetch a list of all recent disputes that the co-ordinator is aware of.
    /// These are disputes which have occurred any time in recent sessions, which may have already concluded.
    RecentDisputes(ResponseChannel<Vec<(SessionIndex, CandidateHash)>>),
    /// Fetch a list of all active disputes that the co-ordinator is aware of.
    /// These disputes are either unconcluded or recently concluded.
    ActiveDisputes(ResponseChannel<Vec<(SessionIndex, CandidateHash)>>),
    /// Get candidate votes for a candidate.
    QueryCandidateVotes(SessionIndex, CandidateHash, ResponseChannel<Option<CandidateVotes>>),
    /// Sign and issue local dispute votes. A value of `true` indicates validity, and `false` invalidity.
    IssueLocalStatement(SessionIndex, CandidateHash, CandidateReceipt, bool),
    /// Determine the highest undisputed block within the given chain, based on where candidates
    /// were included. If even the base block should not be finalized due to a dispute,
    /// then `None` should be returned on the channel.
    ///
    /// The block descriptions begin counting upwards from the block after the given `base_number`. The `base_number`
    /// is typically the number of the last finalized block but may be slightly higher. This block
    /// is inevitably going to be finalized so it is not accounted for by this function.
    DetermineUndisputedChain {
        base_number: BlockNumber,
        block_descriptions: Vec<(BlockHash, SessionIndex, Vec<CandidateHash>)>,
        rx: ResponseSender<Option<(BlockNumber, BlockHash)>>,
    }
}

/// Result of `ImportStatements`.
pub enum ImportStatementsResult {
	/// Import was invalid (candidate was not available)  and the sending peer should get banned.
	InvalidImport,
	/// Import was valid and can be confirmed to peer.
	ValidImport
}
}

Dispute Distribution Message

Messages received by the Dispute Distribution subsystem. This subsystem is responsible of distributing explicit dispute statements.

#![allow(unused)]
fn main() {
enum DisputeDistributionMessage {

  /// Tell dispute distribution to distribute an explicit dispute statement to
  /// validators.
  SendDispute((ValidVote, InvalidVote)),

  /// Ask DisputeDistribution to get votes we don't know about.
  /// Fetched votes will be reported via `DisputeCoordinatorMessage::ImportStatements`
  FetchMissingVotes {
    candidate_hash: CandidateHash,
    session: SessionIndex,
    known_valid_votes: Bitfield,
    known_invalid_votes: Bitfield,
    /// Optional validator to query from. `ValidatorIndex` as in the above
    /// referenced session.
    from_validator: Option<ValidatorIndex>,
  }
}
}

Network Bridge Message

Messages received by the network bridge. This subsystem is invoked by others to manipulate access to the low-level networking code.

#![allow(unused)]
fn main() {
/// Peer-sets handled by the network bridge.
enum PeerSet {
    /// The collation peer-set is used to distribute collations from collators to validators.
    Collation,
    /// The validation peer-set is used to distribute information relevant to parachain
    /// validation among validators. This may include nodes which are not validators,
    /// as some protocols on this peer-set are expected to be gossip.
    Validation,
}

enum NetworkBridgeMessage {
    /// Report a cost or benefit of a peer. Negative values are costs, positive are benefits.
    ReportPeer(PeerId, cost_benefit: i32),
    /// Disconnect a peer from the given peer-set without affecting their reputation.
    DisconnectPeer(PeerId, PeerSet),
    /// Send a message to one or more peers on the validation peerset.
    SendValidationMessage([PeerId], ValidationProtocolV1),
    /// Send a message to one or more peers on the collation peerset.
    SendCollationMessage([PeerId], ValidationProtocolV1),
    /// Send multiple validation messages.
    SendValidationMessages([([PeerId, ValidationProtocolV1])]),
    /// Send multiple collation messages.
    SendCollationMessages([([PeerId, ValidationProtocolV1])]),
    /// Connect to peers who represent the given `validator_ids`.
    ///
    /// Also ask the network to stay connected to these peers at least
    /// until a new request is issued.
    ///
    /// Because it overrides the previous request, it must be ensured
    /// that `validator_ids` include all peers the subsystems
    /// are interested in (per `PeerSet`).
    ///
    /// A caller can learn about validator connections by listening to the
    /// `PeerConnected` events from the network bridge.
    ConnectToValidators {
        /// Ids of the validators to connect to.
        validator_ids: HashSet<AuthorityDiscoveryId>,
        /// The underlying protocol to use for this request.
        peer_set: PeerSet,
        /// Sends back the number of `AuthorityDiscoveryId`s which
        /// authority discovery has failed to resolve.
        failed: oneshot::Sender<usize>,
    },
    /// Inform the distribution subsystems about the new
    /// gossip network topology formed.
    NewGossipTopology {
		/// The session info this gossip topology is concerned with.
		session: SessionIndex,
		/// Our validator index in the session, if any.
		local_index: Option<ValidatorIndex>,
		/// The canonical shuffling of validators for the session.
		canonical_shuffling: Vec<(AuthorityDiscoveryId, ValidatorIndex)>,
		/// The reverse mapping of `canonical_shuffling`: from validator index
		/// to the index in `canonical_shuffling`
		shuffled_indices: Vec<usize>,
    }
}
}

Misbehavior Report

#![allow(unused)]
fn main() {
pub type Misbehavior = generic::Misbehavior<
    CommittedCandidateReceipt,
    CandidateHash,
    ValidatorIndex,
    ValidatorSignature,
>;

mod generic {
    /// Misbehavior: voting more than one way on candidate validity.
    ///
    /// Since there are three possible ways to vote, a double vote is possible in
    /// three possible combinations (unordered)
    pub enum ValidityDoubleVote<Candidate, Digest, Signature> {
        /// Implicit vote by issuing and explicitly voting validity.
        IssuedAndValidity((Candidate, Signature), (Digest, Signature)),
        /// Implicit vote by issuing and explicitly voting invalidity
        IssuedAndInvalidity((Candidate, Signature), (Digest, Signature)),
        /// Direct votes for validity and invalidity
        ValidityAndInvalidity(Candidate, Signature, Signature),
    }

    /// Misbehavior: multiple signatures on same statement.
    pub enum DoubleSign<Candidate, Digest, Signature> {
        /// On candidate.
        Candidate(Candidate, Signature, Signature),
        /// On validity.
        Validity(Digest, Signature, Signature),
        /// On invalidity.
        Invalidity(Digest, Signature, Signature),
    }

    /// Misbehavior: declaring multiple candidates.
    pub struct MultipleCandidates<Candidate, Signature> {
        /// The first candidate seen.
        pub first: (Candidate, Signature),
        /// The second candidate seen.
        pub second: (Candidate, Signature),
    }

    /// Misbehavior: submitted statement for wrong group.
    pub struct UnauthorizedStatement<Candidate, Digest, AuthorityId, Signature> {
        /// A signed statement which was submitted without proper authority.
        pub statement: SignedStatement<Candidate, Digest, AuthorityId, Signature>,
    }

    pub enum Misbehavior<Candidate, Digest, AuthorityId, Signature> {
        /// Voted invalid and valid on validity.
        ValidityDoubleVote(ValidityDoubleVote<Candidate, Digest, Signature>),
        /// Submitted multiple candidates.
        MultipleCandidates(MultipleCandidates<Candidate, Signature>),
        /// Submitted a message that was unauthorized.
        UnauthorizedStatement(UnauthorizedStatement<Candidate, Digest, AuthorityId, Signature>),
        /// Submitted two valid signatures for the same message.
        DoubleSign(DoubleSign<Candidate, Digest, Signature>),
    }
}
}

PoV Distribution Message

This is a network protocol that receives messages of type PoVDistributionV1Message.

#![allow(unused)]
fn main() {
enum PoVDistributionMessage {
    /// Fetch a PoV from the network.
    ///
    /// This `CandidateDescriptor` should correspond to a candidate seconded under the provided
    /// relay-parent hash.
    FetchPoV(Hash, CandidateDescriptor, ResponseChannel<PoV>),
    /// Distribute a PoV for the given relay-parent and CandidateDescriptor.
    /// The PoV should correctly hash to the PoV hash mentioned in the CandidateDescriptor
    DistributePoV(Hash, CandidateDescriptor, PoV),
    /// An update from the network bridge.
    NetworkBridgeUpdate(NetworkBridgeEvent<PoVDistributionV1Message>),
}
}

Provisioner Message

#![allow(unused)]
fn main() {
/// This data becomes intrinsics or extrinsics which should be included in a future relay chain block.
enum ProvisionableData {
  /// This bitfield indicates the availability of various candidate blocks.
  Bitfield(Hash, SignedAvailabilityBitfield),
  /// The Candidate Backing subsystem believes that this candidate is valid, pending availability.
  BackedCandidate(CandidateReceipt),
  /// Misbehavior reports are self-contained proofs of validator misbehavior.
  MisbehaviorReport(Hash, MisbehaviorReport),
  /// Disputes trigger a broad dispute resolution process.
  Dispute(Hash, Signature),
}

/// Message to the Provisioner.
///
/// In all cases, the Hash is that of the relay parent.
enum ProvisionerMessage {
  /// This message allows external subsystems to request current inherent data that could be used for
  /// advancing the state of parachain consensus in a block building upon the given hash.
  ///
  /// If called at different points in time, this may give different results.
  RequestInherentData(Hash, oneshot::Sender<ParaInherentData>),
  /// This data should become part of a relay chain block
  ProvisionableData(ProvisionableData),
}
}

Runtime API Message

The Runtime API subsystem is responsible for providing an interface to the state of the chain's runtime.

This is fueled by an auxiliary type encapsulating all request types defined in the Runtime API section of the guide.

#![allow(unused)]
fn main() {
enum RuntimeApiRequest {
    /// Get the version of the runtime API at the given parent hash, if any.
    Version(ResponseChannel<u32>),
    /// Get the current validator set.
    Validators(ResponseChannel<Vec<ValidatorId>>),
    /// Get the validator groups and rotation info.
    ValidatorGroups(ResponseChannel<(Vec<Vec<ValidatorIndex>>, GroupRotationInfo)>),
    /// Get information about all availability cores.
    AvailabilityCores(ResponseChannel<Vec<CoreState>>),
    /// with the given occupied core assumption.
    PersistedValidationData(
        ParaId,
        OccupiedCoreAssumption,
        ResponseChannel<Option<PersistedValidationData>>,
    ),
    /// Sends back `true` if the commitments pass all acceptance criteria checks.
    CheckValidationOutputs(
        ParaId,
        CandidateCommitments,
        RuntimeApiSender<bool>,
    ),
    /// Get the session index for children of the block. This can be used to construct a signing
    /// context.
    SessionIndexForChild(ResponseChannel<SessionIndex>),
    /// Get the validation code for a specific para, using the given occupied core assumption.
    ValidationCode(ParaId, OccupiedCoreAssumption, ResponseChannel<Option<ValidationCode>>),
    /// Get validation code by its hash, either past, current or future code can be returned,
    /// as long as state is still available.
    ValidationCodeByHash(ValidationCodeHash, RuntimeApiSender<Option<ValidationCode>>),
    /// Get a committed candidate receipt for all candidates pending availability.
    CandidatePendingAvailability(ParaId, ResponseChannel<Option<CommittedCandidateReceipt>>),
    /// Get all events concerning candidates in the last block.
    CandidateEvents(ResponseChannel<Vec<CandidateEvent>>),
    /// Get the session info for the given session, if stored.
    SessionInfo(SessionIndex, ResponseChannel<Option<SessionInfo>>),
    /// Get all the pending inbound messages in the downward message queue for a para.
    DmqContents(ParaId, ResponseChannel<Vec<InboundDownwardMessage<BlockNumber>>>),
    /// Get the contents of all channels addressed to the given recipient. Channels that have no
    /// messages in them are also included.
    InboundHrmpChannelsContents(ParaId, ResponseChannel<BTreeMap<ParaId, Vec<InboundHrmpMessage<BlockNumber>>>>),
    /// Get information about the BABE epoch this block was produced in.
    BabeEpoch(ResponseChannel<BabeEpoch>),
}

enum RuntimeApiMessage {
    /// Make a request of the runtime API against the post-state of the given relay-parent.
    Request(Hash, RuntimeApiRequest),
    /// Get the version of the runtime API at the given parent hash, if any.
    Version(Hash, ResponseChannel<Option<u32>>)
}
}

Statement Distribution Message

The Statement Distribution subsystem distributes signed statements and candidates from validators to other validators. It does this by distributing full statements, which embed the candidate receipt, as opposed to compact statements which don't. It receives updates from the network bridge and signed statements to share with other validators.

This is a network protocol that receives messages of type StatementDistributionV1Message.

#![allow(unused)]
fn main() {
enum StatementDistributionMessage {
    /// An update from the network bridge.
    NetworkBridgeUpdate(NetworkBridgeEvent<StatementDistributionV1Message>),
    /// We have validated a candidate and want to share our judgment with our peers.
    /// The hash is the relay parent.
    ///
    /// The statement distribution subsystem assumes that the statement should be correctly
    /// signed.
    Share(Hash, SignedFullStatementWithPVD),
}
}

Validation Request Type

Various modules request that the Candidate Validation subsystem validate a block with this message. It returns ValidationOutputs for successful validation.

#![allow(unused)]

fn main() {
/// The outcome of the candidate-validation's PVF pre-check request.
pub enum PreCheckOutcome {
    /// The PVF has been compiled successfully within the given constraints.
    Valid,
    /// The PVF could not be compiled. This variant is used when the candidate-validation subsystem
    /// can be sure that the PVF is invalid. To give a couple of examples: a PVF that cannot be
    /// decompressed or that does not represent a structurally valid WebAssembly file.
    Invalid,
    /// This variant is used when the PVF cannot be compiled but for other reasons that are not
    /// included into [`PreCheckOutcome::Invalid`]. This variant can indicate that the PVF in
    /// question is invalid, however it is not necessary that PVF that received this judgement
    /// is invalid.
    ///
    /// For example, if during compilation the preparation worker was killed we cannot be sure why
    /// it happened: because the PVF was malicious made the worker to use too much memory or its
    /// because the host machine is under severe memory pressure and it decided to kill the worker.
    Failed,
}

/// Result of the validation of the candidate.
enum ValidationResult {
    /// Candidate is valid, and here are the outputs and the validation data used to form inputs.
    /// In practice, this should be a shared type so that validation caching can be done.
    Valid(CandidateCommitments, PersistedValidationData),
    /// Candidate is invalid.
    Invalid,
}

const BACKING_EXECUTION_TIMEOUT: Duration = 2 seconds;
const APPROVAL_EXECUTION_TIMEOUT: Duration = 6 seconds;

/// Messages received by the Validation subsystem.
///
/// ## Validation Requests
///
/// Validation requests made to the subsystem should return an error only on internal error.
/// Otherwise, they should return either `Ok(ValidationResult::Valid(_))`
/// or `Ok(ValidationResult::Invalid)`.
#[derive(Debug)]
pub enum CandidateValidationMessage {
    /// Validate a candidate with provided parameters using relay-chain state.
    ///
    /// This will implicitly attempt to gather the `PersistedValidationData` and `ValidationCode`
    /// from the runtime API of the chain, based on the `relay_parent`
    /// of the `CandidateDescriptor`.
    ///
    /// This will also perform checking of validation outputs against the acceptance criteria.
    ///
    /// If there is no state available which can provide this data or the core for
    /// the para is not free at the relay-parent, an error is returned.
    ValidateFromChainState(
        CandidateDescriptor,
        Arc<PoV>,
        Duration, // Execution timeout.
        oneshot::Sender<Result<ValidationResult, ValidationFailed>>,
    ),
    /// Validate a candidate with provided, exhaustive parameters for validation.
    ///
    /// Explicitly provide the `PersistedValidationData` and `ValidationCode` so this can do full
    /// validation without needing to access the state of the relay-chain.
    ///
    /// This request doesn't involve acceptance criteria checking, therefore only useful for the
    /// cases where the validity of the candidate is established. This is the case for the typical
    /// use-case: approval checkers would use this request relying on the full prior checks
    /// performed by the relay-chain.
    ValidateFromExhaustive(
        PersistedValidationData,
        ValidationCode,
        CandidateDescriptor,
        Arc<PoV>,
        Duration, // Execution timeout.
        oneshot::Sender<Result<ValidationResult, ValidationFailed>>,
    ),
    /// Try to compile the given validation code and send back
    /// the outcome.
    ///
    /// The validation code is specified by the hash and will be queried from the runtime API at the
    /// given relay-parent.
    PreCheck(
        // Relay-parent
        Hash,
        ValidationCodeHash,
        oneshot::Sender<PreCheckOutcome>,
    ),
}
}

PVF Pre-checker Message

Currently, the PVF pre-checker subsystem receives no specific messages.

#![allow(unused)]
fn main() {
/// Non-instantiable message type
pub enum PvfCheckerMessage { }
}

Runtime

Types used within the runtime exclusively and pervasively.

Host Configuration

The internal-to-runtime configuration of the parachain host. This is expected to be altered only by governance procedures.

#![allow(unused)]
fn main() {
struct HostConfiguration {
	/// The minimum period, in blocks, between which parachains can update their validation code.
	pub validation_upgrade_cooldown: BlockNumber,
	/// The delay, in blocks, before a validation upgrade is applied.
	pub validation_upgrade_delay: BlockNumber,
	/// How long to keep code on-chain, in blocks. This should be sufficiently long that disputes
	/// have concluded.
	pub code_retention_period: BlockNumber,
	/// The maximum validation code size, in bytes.
	pub max_code_size: u32,
	/// The maximum head-data size, in bytes.
	pub max_head_data_size: u32,
	/// The amount of availability cores to dedicate to parathreads (on-demand parachains).
	pub parathread_cores: u32,
	/// The number of retries that a parathread (on-demand parachain) author has to submit their block.
	pub parathread_retries: u32,
	/// How often parachain groups should be rotated across parachains.
	pub group_rotation_frequency: BlockNumber,
	/// The availability period, in blocks, for parachains. This is the amount of blocks
	/// after inclusion that validators have to make the block available and signal its availability to
	/// the chain. Must be at least 1.
	pub chain_availability_period: BlockNumber,
	/// The availability period, in blocks, for parathreads (on-demand parachains). Same as the `chain_availability_period`,
	/// but a differing timeout due to differing requirements. Must be at least 1.
	pub thread_availability_period: BlockNumber,
	/// The amount of blocks ahead to schedule on-demand parachains.
	pub scheduling_lookahead: u32,
	/// The maximum number of validators to have per core. `None` means no maximum.
	pub max_validators_per_core: Option<u32>,
	/// The maximum number of validators to use for parachains, in total. `None` means no maximum.
	pub max_validators: Option<u32>,
	/// The amount of sessions to keep for disputes.
	pub dispute_period: SessionIndex,
	/// How long after dispute conclusion to accept statements.
	pub dispute_post_conclusion_acceptance_period: BlockNumber,
	/// The maximum number of dispute spam slots
	pub dispute_max_spam_slots: u32,
	/// The amount of consensus slots that must pass between submitting an assignment and
	/// submitting an approval vote before a validator is considered a no-show.
	/// Must be at least 1.
	pub no_show_slots: u32,
	/// The number of delay tranches in total.
	pub n_delay_tranches: u32,
	/// The width of the zeroth delay tranche for approval assignments. This many delay tranches
	/// beyond 0 are all consolidated to form a wide 0 tranche.
	pub zeroth_delay_tranche_width: u32,
	/// The number of validators needed to approve a block.
	pub needed_approvals: u32,
	/// The number of samples to do of the RelayVRFModulo approval assignment criterion.
	pub relay_vrf_modulo_samples: u32,
	/// Total number of individual messages allowed in the parachain -> relay-chain message queue.
	pub max_upward_queue_count: u32,
	/// Total size of messages allowed in the parachain -> relay-chain message queue before which
	/// no further messages may be added to it. If it exceeds this then the queue may contain only
	/// a single message.
	pub max_upward_queue_size: u32,
	/// The maximum size of an upward message that can be sent by a candidate.
	///
	/// This parameter affects the upper bound of size of `CandidateCommitments`.
	pub max_upward_message_size: u32,
	/// The maximum number of messages that a candidate can contain.
	///
	/// This parameter affects the upper bound of size of `CandidateCommitments`.
	pub max_upward_message_num_per_candidate: u32,
	/// The maximum size of a message that can be put in a downward message queue.
	///
	/// Since we require receiving at least one DMP message the obvious upper bound of the size is
	/// the PoV size. Of course, there is a lot of other different things that a parachain may
	/// decide to do with its PoV so this value in practice will be picked as a fraction of the PoV
	/// size.
	pub max_downward_message_size: u32,
	/// The deposit that the sender should provide for opening an HRMP channel.
	pub hrmp_sender_deposit: u32,
	/// The deposit that the recipient should provide for accepting opening an HRMP channel.
	pub hrmp_recipient_deposit: u32,
	/// The maximum number of messages allowed in an HRMP channel at once.
	pub hrmp_channel_max_capacity: u32,
	/// The maximum total size of messages in bytes allowed in an HRMP channel at once.
	pub hrmp_channel_max_total_size: u32,
	/// The maximum number of inbound HRMP channels a parachain is allowed to accept.
	pub hrmp_max_parachain_inbound_channels: u32,
	/// The maximum number of inbound HRMP channels a parathread (on-demand parachain) is allowed to accept.
	pub hrmp_max_parathread_inbound_channels: u32,
	/// The maximum size of a message that could ever be put into an HRMP channel.
	///
	/// This parameter affects the upper bound of size of `CandidateCommitments`.
	pub hrmp_channel_max_message_size: u32,
	/// The maximum number of outbound HRMP channels a parachain is allowed to open.
	pub hrmp_max_parachain_outbound_channels: u32,
	/// The maximum number of outbound HRMP channels a parathread (on-demand parachain) is allowed to open.
	pub hrmp_max_parathread_outbound_channels: u32,
	/// The maximum number of outbound HRMP messages can be sent by a candidate.
	///
	/// This parameter affects the upper bound of size of `CandidateCommitments`.
	pub hrmp_max_message_num_per_candidate: u32,
}
}

ParaInherentData

Inherent data passed to a runtime entry-point for the advancement of parachain consensus.

This contains 4 pieces of data:

  1. Bitfields
  2. BackedCandidates
  3. MultiDisputeStatementSet
  4. Header
#![allow(unused)]
fn main() {
struct ParaInherentData {
	bitfields: Bitfields,
	backed_candidates: BackedCandidates,
	dispute_statements: MultiDisputeStatementSet,
	parent_header: Header
}
}

Message types

Types of messages that are passed between parachains and the relay chain: UMP, DMP, XCMP.

There is also HRMP (Horizontally Relay-routed Message Passing) which provides the same functionality although with smaller scalability potential.

Vertical Message Passing

Types required for message passing between the relay-chain and a parachain.

Actual contents of the messages is specified by the XCM standard.

/// A message sent from a parachain to the relay-chain.
type UpwardMessage = Vec<u8>;

/// A message sent from the relay-chain down to a parachain.
///
/// The size of the message is limited by the `config.max_downward_message_size`
/// parameter.
type DownwardMessage = Vec<u8>;

/// This struct extends `DownwardMessage` by adding the relay-chain block number when the message was
/// enqueued in the downward message queue.
struct InboundDownwardMessage {
	/// The block number at which this messages was put into the downward message queue.
	pub sent_at: BlockNumber,
	/// The actual downward message to processes.
	pub msg: DownwardMessage,
}

Horizontal Message Passing

HrmpChannelId

A type that uniquely identifies an HRMP channel. An HRMP channel is established between two paras. In text, we use the notation (A, B) to specify a channel between A and B. The channels are unidirectional, meaning that (A, B) and (B, A) refer to different channels. The convention is that we use the first item tuple for the sender and the second for the recipient. Only one channel is allowed between two participants in one direction, i.e. there cannot be 2 different channels identified by (A, B).

struct HrmpChannelId {
    sender: ParaId,
    recipient: ParaId,
}

Horizontal Message

This is a message sent from a parachain to another parachain that travels through the relay chain. This message ends up in the recipient's mailbox. A size of a horizontal message is defined by its data payload.

struct OutboundHrmpMessage {
	/// The para that will get this message in its downward message queue.
	pub recipient: ParaId,
	/// The message payload.
	pub data: Vec<u8>,
}

struct InboundHrmpMessage {
	/// The block number at which this message was sent.
	/// Specifically, it is the block number at which the candidate that sends this message was
	/// enacted.
	pub sent_at: BlockNumber,
	/// The message payload.
	pub data: Vec<u8>,
}

Network Types

These types are those that are actually sent over the network to subsystems.

Universal Types

#![allow(unused)]
fn main() {
type RequestId = u64;
type ProtocolVersion = u32;
struct PeerId(...); // opaque, unique identifier of a peer.
struct View {
	// Up to `N` (5?) chain heads.
	heads: Vec<Hash>,
	// The number of the finalized block.
	finalized_number: BlockNumber,
}

enum ObservedRole {
	Full,
	Light,
}
}

V1 Network Subsystem Message Types

Approval Distribution V1

#![allow(unused)]
fn main() {
enum ApprovalDistributionV1Message {
	/// Assignments for candidates in recent, unfinalized blocks.
	///
	/// The u32 is the claimed index of the candidate this assignment corresponds to. Actually checking the assignment
	/// may yield a different result.
	Assignments(Vec<(IndirectAssignmentCert, u32)>),
	/// Approvals for candidates in some recent, unfinalized block.
	Approvals(Vec<IndirectSignedApprovalVote>),
}
}

Availability Distribution V1

#![allow(unused)]
fn main() {
enum AvailabilityDistributionV1Message {
	/// An erasure chunk for a given candidate hash.
	Chunk(CandidateHash, ErasureChunk),
}
}

Availability Recovery V1

#![allow(unused)]
fn main() {
enum AvailabilityRecoveryV1Message {
	/// Request a chunk for a given candidate hash and validator index.
	RequestChunk(RequestId, CandidateHash, ValidatorIndex),
	/// Respond with chunk for a given candidate hash and validator index.
	/// The response may be `None` if the requestee does not have the chunk.
	Chunk(RequestId, Option<ErasureChunk>),
	/// Request the full data for a given candidate hash.
	RequestFullData(RequestId, CandidateHash),
	/// Respond with data for a given candidate hash and validator index.
	/// The response may be `None` if the requestee does not have the data.
	FullData(RequestId, Option<AvailableData>),

}
}

Bitfield Distribution V1

#![allow(unused)]
fn main() {
enum BitfieldDistributionV1Message {
	/// A signed availability bitfield for a given relay-parent hash.
	Bitfield(Hash, SignedAvailabilityBitfield),
}
}

PoV Distribution V1

#![allow(unused)]
fn main() {
enum PoVDistributionV1Message {
	/// Notification that we are awaiting the given PoVs (by hash) against a
	/// specific relay-parent hash.
	Awaiting(Hash, Vec<Hash>),
	/// Notification of an awaited PoV, in a given relay-parent context.
	/// (`relay_parent`, `pov_hash`, `pov`)
	SendPoV(Hash, Hash, PoV),
}
}

Statement Distribution V1

#![allow(unused)]
fn main() {
enum StatementDistributionV1Message {
	/// A signed full statement under a given relay-parent.
	Statement(Hash, SignedFullStatement)
}
}

Collator Protocol V1

#![allow(unused)]
fn main() {
enum CollatorProtocolV1Message {
	/// Declare the intent to advertise collations under a collator ID and `Para`, attaching a
	/// signature of the `PeerId` of the node using the given collator ID key.
	Declare(CollatorId, ParaId, CollatorSignature),
	/// Advertise a collation to a validator. Can only be sent once the peer has
	/// declared that they are a collator with given ID.
	AdvertiseCollation(Hash),
	/// A collation sent to a validator was seconded.
	CollationSeconded(SignedFullStatement),
}
}

V1 Wire Protocols

Validation V1

These are the messages for the protocol on the validation peer-set.

#![allow(unused)]
fn main() {
enum ValidationProtocolV1 {
	ApprovalDistribution(ApprovalDistributionV1Message),
	AvailabilityDistribution(AvailabilityDistributionV1Message),
	AvailabilityRecovery(AvailabilityRecoveryV1Message),
	BitfieldDistribution(BitfieldDistributionV1Message),
	PoVDistribution(PoVDistributionV1Message),
	StatementDistribution(StatementDistributionV1Message),
}
}

Collation V1

These are the messages for the protocol on the collation peer-set

#![allow(unused)]
fn main() {
enum CollationProtocolV1 {
	CollatorProtocol(CollatorProtocolV1Message),
}
}

Network Bridge Event

These updates are posted from the Network Bridge Subsystem to other subsystems based on registered listeners.

#![allow(unused)]
fn main() {
struct NewGossipTopology {
	/// The session index this topology corresponds to.
	session: SessionIndex,
	/// The topology itself.
	topology: SessionGridTopology,
	/// The local validator index, if any.
	local_index: Option<ValidatorIndex>,
}

struct SessionGridTopology {
	/// An array mapping validator indices to their indices in the
	/// shuffling itself. This has the same size as the number of validators
	/// in the session.
	shuffled_indices: Vec<usize>,
	/// The canonical shuffling of validators for the session.
	canonical_shuffling: Vec<TopologyPeerInfo>,
}

struct TopologyPeerInfo {
	/// The validator's known peer IDs.
	peer_ids: Vec<PeerId>,
	/// The index of the validator in the discovery keys of the corresponding
	/// `SessionInfo`. This can extend _beyond_ the set of active parachain validators.
	validator_index: ValidatorIndex,
	/// The authority discovery public key of the validator in the corresponding
	/// `SessionInfo`.
	discovery_id: AuthorityDiscoveryId,
}

enum NetworkBridgeEvent<M> {
	/// A peer with given ID is now connected.
	PeerConnected(PeerId, ObservedRole, ProtocolVersion, Option<HashSet<AuthorityDiscoveryId>>),
	/// A peer with given ID is now disconnected.
	PeerDisconnected(PeerId),
	/// Our neighbors in the new gossip topology.
	/// We're not necessarily connected to all of them.
	///
	/// This message is issued only on the validation peer set.
	///
	/// Note, that the distribution subsystems need to handle the last
	/// view update of the newly added gossip peers manually.
	NewGossipTopology(NewGossipTopology),
	/// We received a message from the given peer.
	PeerMessage(PeerId, M),
	/// The given peer has updated its description of its view.
	PeerViewChange(PeerId, View), // guaranteed to come after peer connected event.
	/// We have posted the given view update to all connected peers.
	OurViewChange(View),
}
}

Approval Types

AssignmentId

The public key of a keypair used by a validator for determining assignments to approve included parachain candidates.

AssignmentCert

An AssignmentCert, short for Assignment Certificate, is a piece of data provided by a validator to prove that they have been selected to perform approval checks on an included candidate.

These certificates can be checked in the context of a specific block, candidate, and validator assignment VRF key. The block state will also provide further context about the availability core states at that block.

#![allow(unused)]
fn main() {
enum AssignmentCertKind {
    RelayVRFModulo {
        sample: u32,
    },
    RelayVRFDelay {
        core_index: CoreIndex,
    }
}

struct AssignmentCert {
    // The criterion which is claimed to be met by this cert.
    kind: AssignmentCertKind,
    // The VRF showing the criterion is met.
    vrf: (VRFPreOut, VRFProof),
}
}

TODO: RelayEquivocation cert. Probably can only be broadcast to chains that have handled an equivocation report.

IndirectAssignmentCert

An assignment cert which refers to the candidate under which the assignment is relevant by block hash.

#![allow(unused)]
fn main() {
struct IndirectAssignmentCert {
    // A block hash where the candidate appears.
    block_hash: Hash,
    validator: ValidatorIndex,
    cert: AssignmentCert,
}
}

ApprovalVote

A vote of approval on a candidate.

#![allow(unused)]
fn main() {
struct ApprovalVote(Hash);
}

SignedApprovalVote

An approval vote signed with a validator's key. This should be verifiable under the ValidatorId corresponding to the ValidatorIndex of the session, which should be implicit from context.

#![allow(unused)]
fn main() {
struct SignedApprovalVote {
    vote: ApprovalVote,
    validator: ValidatorIndex,
    signature: ValidatorSignature,
}
}

IndirectSignedApprovalVote

A signed approval vote which references the candidate indirectly via the block. If there exists a look-up to the candidate hash from the block hash and candidate index, then this can be transformed into a SignedApprovalVote.

Although this vote references the candidate by a specific block hash and candidate index, the signature is computed on the actual SignedApprovalVote payload.

#![allow(unused)]
fn main() {
struct IndirectSignedApprovalVote {
    // A block hash where the candidate appears.
    block_hash: Hash,
    // The index of the candidate in the list of candidates fully included as-of the block.
    candidate_index: CandidateIndex,
    validator: ValidatorIndex,
    signature: ValidatorSignature,
}
}

CheckedAssignmentCert

An assignment cert which has checked both the VRF and the validity of the implied assignment according to the selection criteria rules of the protocol. This type should be declared in such a way as to be instantiatable only when the checks have actually been done. Fields should be accessible via getters, not direct struct access.

#![allow(unused)]
fn main() {
struct CheckedAssignmentCert {
    cert: AssignmentCert,
    validator: ValidatorIndex,
    relay_block: Hash,
    candidate_hash: Hash,
    delay_tranche: DelayTranche,
}
}

DelayTranche

#![allow(unused)]
fn main() {
type DelayTranche = u32;
}

Disputes

DisputeStatementSet

#![allow(unused)]
fn main() {
/// A set of statements about a specific candidate.
struct DisputeStatementSet {
    candidate_hash: CandidateHash,
    session: SessionIndex,
    statements: Vec<(DisputeStatement, ValidatorIndex, ValidatorSignature)>,
}
}

DisputeStatement

#![allow(unused)]
fn main() {
/// A statement about a candidate, to be used within some dispute resolution process.
///
/// Statements are either in favor of the candidate's validity or against it.
enum DisputeStatement {
    /// A valid statement, of the given kind
    Valid(ValidDisputeStatementKind),
    /// An invalid statement, of the given kind.
    Invalid(InvalidDisputeStatementKind),
}

}

Dispute Statement Kinds

Kinds of dispute statements. Each of these can be combined with a candidate hash, session index, validator public key, and validator signature to reproduce and check the original statement.

#![allow(unused)]
fn main() {
enum ValidDisputeStatementKind {
    Explicit,
    BackingSeconded(Hash),
    BackingValid(Hash),
    ApprovalChecking,
}

enum InvalidDisputeStatementKind {
    Explicit,
}
}

ExplicitDisputeStatement

#![allow(unused)]
fn main() {
struct ExplicitDisputeStatement {
    valid: bool,
    candidate_hash: CandidateHash,
    session: SessionIndex,
}
}

MultiDisputeStatementSet

Sets of statements for many (zero or more) disputes.

#![allow(unused)]
fn main() {
type MultiDisputeStatementSet = Vec<DisputeStatementSet>;
}

DisputeState

#![allow(unused)]
fn main() {
struct DisputeState {
    validators_for: Bitfield, // one bit per validator.
    validators_against: Bitfield, // one bit per validator.
    start: BlockNumber,
    concluded_at: Option<BlockNumber>,
}
}

ScrapedOnChainVotes

#![allow(unused)]
fn main() {
/// Type for transcending recorded on-chain
/// dispute relevant votes and conclusions to
/// the off-chain `DisputesCoordinator`.
struct ScrapedOnChainVotes {
    /// The session index at which the block was included.
    session: SessionIndex,
    /// The backing and seconding validity attestations for all candidates, provigind the full candidate receipt.
    backing_validators_per_candidate: Vec<(CandidateReceipt<H>, Vec<(ValidatorIndex, ValidityAttestation)>)>
    /// Set of concluded disputes that were recorded
    /// on chain within the inherent.
    disputes: MultiDisputeStatementSet,
}
}

PVF Pre-checking types

PvfCheckStatement

⚠️ This type was added in v2.

One of the main units of information on which PVF pre-checking voting is build is the PvfCheckStatement.

This is a statement by the validator who ran the pre-checking process for a PVF. A PVF is identified by the ValidationCodeHash.

The statement is valid only during a single session, specified in the session_index.

#![allow(unused)]
fn main() {
struct PvfCheckStatement {
    /// `true` if the subject passed pre-checking and `false` otherwise.
    pub accept: bool,
    /// The validation code hash that was checked.
    pub subject: ValidationCodeHash,
    /// The index of a session during which this statement is considered valid.
    pub session_index: SessionIndex,
    /// The index of the validator from which this statement originates.
    pub validator_index: ValidatorIndex,
}
}

Glossary

Here you can find definitions of a bunch of jargon, usually specific to the Polkadot project.

  • Approval Checker: A validator who randomly self-selects so to perform validity checks on a parablock which is pending approval.
  • BABE: (Blind Assignment for Blockchain Extension). The algorithm validators use to safely extend the Relay Chain. See the Polkadot wiki for more information.
  • Backable Candidate: A Parachain Candidate which is backed by a majority of validators assigned to a given parachain.
  • Backed Candidate: A Backable Candidate noted in a relay-chain block
  • Backing: A set of statements proving that a Parachain Candidate is backable.
  • Collator: A node who generates Proofs-of-Validity (PoV) for blocks of a specific parachain.
  • DMP: (Downward Message Passing). Message passing from the relay-chain to a parachain. Also there is a runtime parachains module with the same name.
  • DMQ: (Downward Message Queue). A message queue for messages from the relay-chain down to a parachain. A parachain has exactly one downward message queue.
  • Extrinsic: An element of a relay-chain block which triggers a specific entry-point of a runtime module with given arguments.
  • GRANDPA: (Ghost-based Recursive ANcestor Deriving Prefix Agreement). The algorithm validators use to guarantee finality of the Relay Chain.
  • HRMP: (Horizontally Relay-routed Message Passing). A mechanism for message passing between parachains (hence horizontal) that leverages the relay-chain storage. Predates XCMP. Also there is a runtime parachains module with the same name.
  • Inclusion Pipeline: The set of steps taken to carry a Parachain Candidate from authoring, to backing, to availability and full inclusion in an active fork of its parachain.
  • Module: A component of the Runtime logic, encapsulating storage, routines, and entry-points.
  • Module Entry Point: A recipient of new information presented to the Runtime. This may trigger routines.
  • Module Routine: A piece of code executed within a module by block initialization, closing, or upon an entry point being triggered. This may execute computation, and read or write storage.
  • MQC: (Message Queue Chain). A cryptographic data structure that resembles an append-only linked list which doesn't store original values but only their hashes. The whole structure is described by a single hash, referred as a "head". When a value is appended, it's contents hashed with the previous head creating a hash that becomes a new head.
  • Node: A participant in the Polkadot network, who follows the protocols of communication and connection to other nodes. Nodes form a peer-to-peer network topology without a central authority.
  • Parachain Candidate, or Candidate: A proposed block for inclusion into a parachain.
  • Parablock: A block in a parachain.
  • Parachain: A constituent chain secured by the Relay Chain's validators.
  • Parachain Validators: A subset of validators assigned during a period of time to back candidates for a specific parachain
  • On-demand parachain: A parachain which is scheduled on a pay-as-you-go basis.
  • Lease holding parachain: A parachain possessing an active slot lease. The lease holder is assigned a single availability core for the duration of the lease, granting consistent blockspace scheduling at the rate 1 parablock per relay block.
  • PDK (Parachain Development Kit): A toolset that allows one to develop a parachain. Cumulus is a PDK.
  • Preimage: In our context, if H(X) = Y where H is a hash function and Y is the hash, then X is the hash preimage.
  • Proof-of-Validity (PoV): A stateless-client proof that a parachain candidate is valid, with respect to some validation function.
  • PVF: Parachain Validation Function. The validation code that is run by validators on parachains.
  • PVF Prechecking: This is the process of initially checking the PVF when it is first added. We attempt preparation of the PVF and make sure it succeeds within a given timeout, plus some additional checks.
  • PVF Preparation: This is the process of preparing the WASM blob and includes both prevalidation and compilation. As there is no prevalidation right now, preparation just consists of compilation.
  • Relay Parent: A block in the relay chain, referred to in a context where work is being done in the context of the state at this block.
  • Runtime: The relay-chain state machine.
  • Runtime Module: See Module.
  • Runtime API: A means for the node-side behavior to access structured information based on the state of a fork of the blockchain.
  • Subsystem: A long-running task which is responsible for carrying out a particular category of work.
  • UMP: (Upward Message Passing) A vertical message passing mechanism from a parachain to the relay chain.
  • Validator: Specially-selected node in the network who is responsible for validating parachain blocks and issuing attestations about their validity.
  • Validation Function: A piece of Wasm code that describes the state-transition function of a parachain.
  • VMP: (Vertical Message Passing) A family of mechanisms that are responsible for message exchange between the relay chain and parachains.
  • XCMP: (Cross-Chain Message Passing) A type of horizontal message passing (i.e. between parachains) that allows secure message passing directly between parachains and has minimal resource requirements from the relay chain, thus highly scalable.

See Also

Also of use is the Substrate Glossary.

Further Reading