ImperialViolet

Real World Crypto 2013 (13 Jan 2013)

(These are my notes for a talk that I gave last week at Real World Crypto. The premise of the conference is that it brings together theoretical cryptographers and practitioners. So this talk is aimed at theoretical cryptographers but it's fairly simple as I don't have anything complex worth saying to real cryptographers! Slides for other talks are linked from the program page and Rogaway's are relevant to this talk.

Note that this isn't a transcript: I actually say more words than this, but it contains the main points.)

Hello all.

For those who don't know me, I'm Adam Langley. I work at Google, mostly on our serving side HTTPS infrastructure these days. But I also still do some work on Chrome's SSL stack from time to time.

When I was asked to come up with a title of a talk I picked “Things that bit us, things we fixed and things that are waiting in the grass” with reference to HTTPS. Partly because that's what I know about but also because HTTPS is the only place where most people knowingly interact with public crypto, so it's one of the few examples of real world crypto at really large, really messy scales.

I could also have titled this talk “Know your enemy” because your real enemy is not trying to achieve an advantage with greater than negligible probability. As I hope to convince you, your enemy is me. I am the idiot who will mess up the implementation of your lovely cryptosystem. Your 128-bit security level is worthless in the face of my stupidity and so I'm here to seek your help in making my life easier and, therefore, everything more secure.

But before we get into anything crypto related, I'd like to set the scene. By and large, transport security on the Internet is doing OK because few people bother to attack it. If you want to steal banking credentials, you can get malware toolkits that will work on a large enough fraction of machines to be useful - e.g. using the Java exploit made public this week. If you want to steal passwords, you can SQL inject the site and reverse the hashes (if the site even bothered to hash), or just use the Ruby on Rails exploit made public this week. Given the level of password reuse, it doesn't even have to be the same site! Economically, attacking the transport doesn't make sense for many attackers and so they don't do it.

If you do want to attack the transport, by far the best methods are SSL stripping, mixed scripting and insecure cookie vulnerabilities.

SSL stripping means that, when the user types in example.com, since the default scheme is unencrypted, the browser makes an insecure request. The attacker can simply answer that request and proxy the entire site while removing any attempts to upgrade to HTTPS. In the majority of cases, the user will never notice.

Mixed scripting results from pages that source Javascript, CSS or plugins from an HTTP URL. The request is made insecurely but the response is trusted with the same authority as the HTTPS origin that requested them. On sites that serve both HTTP and HTTPS, mixed scripting is endemic.

Lastly, cookies are secrets that the client sends to authenticate requests. If, when creating a cookie, the server doesn't set the secure flag, the client will also send the same cookie over unencrypted connections. Since forgetting to set that flag doesn't break anything in normal operation, it happens fairly often.

Oh, and all that is assuming that any transport security is used at all. HTTP is still much more common than HTTPS.

I've breezed over those issues, but the important point is that none of them involve crypto, they're all systems issues. On the whole, the crypto is doing great in comparison to everything else!

I'll cycle back to those issues towards the end but I wanted to give some perspective as I'll be talking about the crypto a lot, since this is Real World Crypto, but crypto is only a small part of web security.

When writing this talk I sat down and made a list of the cryptographic issues that have bitten HTTPS over the past few years and tried to group in order to make points that I think would be useful for the research community. I didn't end up using them all, so this is hardly a exhaustive list and I used some things twice, but here they are:

Go beyond a paper

  1. CRIME (compression leaks.)
  2. BEAST (CBC vs adaptive, chosen plaintext.)

The first group of issues I called ‘go beyond a paper’:

The CRIME attack resulted from the attacker having partial, chosen plaintext abilities on a channel which was performing compression. This applied to both TLS, which supports compression, and SPDY, an HTTP replacement that runs over TLS that applied compression internally. Most, major HTTPS sites don't support TLS compression so SPDY was the bigger hole, but SPDY is a relatively new protocol and the idea that compression can leak information isn't. I can easily find it going back ten years (“Compression and Information Leakage of Plaintext”, Kelsey, 2002) and here's what I said on the SPDY mailing list before the CRIME work:

With a good model of zlib, I think you could extract a ~40 byte cookie with ~13K requests. That's a practical attack and would make a great paper if someone has the time.

Of course, I never had the time but we did start working on a better compression scheme to solve the issue.

But SPDY was done by smart networking folks. Since the compression leak issues weren't known widely enough at the time, they picked gzip as a seemingly reasonable compression algorithm. It wasn't clearly stupid to compose TLS and gzip and yet it blew up when they did so. Had the Kelsey paper instead been a splashy, public demonstration, as CRIME was, then it's possible that the idea would have sunk into the collective memory to the point where simply using gzip wouldn't have seemed so reasonable. As it is, we're now using a horrible gzip hack to segment cookies and other sensitive headers from the attacker controlled values, although the replacement compression is mostly done, pending SPDY 4.

So the first idea in this section is that it's OK to make a splash in order to get some new idea into the minds of the developer community in general.

Somewhat similarly, the origins of the BEAST attack dates back to at least 2002 in the context of SSH (Möller, 2002).

In this case, Rizzo and Duong contacted the major SSL stacks prior to their work going public with a clear case that it was worth looking into. This, at least from my point of view, is very welcome! Please do this if you can. You can contact me if nothing else and I can rope in all the other usual suspects. Please set a clear go-public date and allow us to share the facts of the matter with other major vendors.

In the case of BEAST, this produced a rare example of major browsers breaking things together for the common good. The breakage wasn't trivial: we took out Disneyland Tokyo's online ticking system amongst many others, but we got it deployed.

So the second point of this group is that you should consider treating it like a security disclosure if it's warranted. We don't bite!

The world often sucks

  1. Hash collisions in MD5
  2. BEAST
  3. Downgrade attacks

The second group I put together under the title ‘the world often sucks’.

I'm afraid that sometimes it takes a while to make changes even if we are given a very clear demonstration of the issue. After a very clear demo of MD5 collisions causing vulnerabilities in certificate issuance in 2008 it still took years to remove support for it. Other remedial actions were taken: public CAs were required to use random serial numbers and stopped using MD5 for new certificates. But it wasn't until early 2012 that Chrome removed support for MD5 in signatures. Sadly, many MITM proxies still used MD5 and, despite giving them lots of notice, they didn't do anything and we broke quite a few people with that change.

Of course, later in the year it was found that Flame broke Microsoft's code signing with an MD5 collision. Maybe that'll be enough to convince people to move.

But the point of this section is to encourage people to think about workarounds because the ‘right fix’ often isn't feasible. For hash collisions we have randomised serial numbers and although we might like everyone to be using SHA-256 or SHA-3, realistically those randomised serial numbers are going to be buttressing SHA-1 for quite some time to come.

Marc Stevens talked earlier about how to detect SHA-1 collisions given only one of the colliding messages. That's great! That's absolutely something that we can put into certificate validation now and hopefully prevent problems in the future.

In relation to BEAST: I mentioned that the core weakness had been known since 2002 and, because of that, there was even a workaround in OpenSSL for it: empty fragments. By sending an empty CBC fragment before each real record, the MAC would effectively randomise the IV. However, empty fragments caused compatibility issues because some SSL stacks returned a zero length when encountering them and higher layers of the code took this to be EOF. Because of that, the workaround was never widely enabled.

In the course of discussing BEAST another workaround was proposed: 1/n-1 record splitting. Rather than putting an empty fragment before each record, include a single byte of the plaintext. The protection isn't quite as good, but it solves the EOF problem. Some servers still broke because they assumed that the complete HTTP request would come in a single read, but the lower rate of problems probably made BEAST mitigation viable.

Lastly, there's the first of our snakes in the grass (problems that are still pending to bite us): SSLv3 downgrade. Since there exist so many broken servers and network middleboxes on the Internet that can't handle TLS version negotiation, in the event of a TLS handshake failure browsers will fall back to SSLv3. However, SSLv3 doesn't support all of the features that TLS does. Most significantly from my point of view, it doesn't support ECDHE, which Google servers use. So a network attacker can trigger a fallback and downgrade a capable client to a non-forward secure ciphersuite.

This is obviously not good. The correct fix is to remove the fallback to SSLv3 of course, but that's sadly not viable right now. Instead, as a workaround, Yngve (formerly of Opera) suggested using the TLS renegotiation extension as a signal that a server is reasonably recent and therefore we shouldn't have performed the fallback.

Numbers from Chrome indicate that a high fraction of the servers that we perform fallback for are renego patched, suggesting that's a bad signal and we should instead create a different one. Although maybe the number of fallbacks is dominated by transient network problems and that's skewing the data. Eric Rescorla has suggested replicating the TLS version negotiation using ciphersuite values. It's something that we will hopefully address in one way or another in 2013.

So that's the point of this group: please consider workarounds because we can't always manage to deploy the ‘right’ fix.

Side-channels are a big deal.

  1. RSA PKCS#1 v1.5 padding oracles ("Million message attack")
  2. CBC padding oracles (Vaudenay's attack)
  3. Timing attacks against RSA CRT.
  4. Side-channel attacks against AES
  5. Side-channel attacks against group operations

This big group is about the fact that side channels are a big deal. Writing constant-time code is a very odd corner of programming and not something that can be easily tested. I have a valgrind hack called ctgrind that allows for some automated testing, but it certainly has its limitations.

But what does constant-time even mean? There's the obvious definition: the CPU runs for the same amount of time, independent of any secret inputs. But CPUs are preemptively multitasked and have frequency scaling and thermal limiting these days so that's not a very useful definition in practice. A more workable definition is that if the code were to be run on an abstract Von Neumann machine, then the trace of memory fetches is identical for all secret inputs. That means that the trace of instructions fetched and data accesses is constant for all secret inputs. That takes care of all the obvious problems.

In practice, it can be useful to relax the definition a little and require only that the set of cache lines accessed for data is constant, rather than the exact addresses. In practice CPUs often fetch whole cache lines at a time and using that fact can lead to speedups at the cost of having to know the cache line length of the CPU.

This model assumes that the individual instructions themselves are constant time. As a research topic it would be interesting to know how variable time CPU instructions affect this. For example, from the Intel optimisation manual:

The latency and throughput of IDIV in Enhanced Intel Core micro-architecture varies with operand sizes and with the number of significant digits of the quotient of the division.

Is this a problem, cryptographically? I don't know. I think multiplies on ARM are variable time too. They are on PowerPC, but that's a much more obscure platform.

As a researcher, what's there to keep in mind with respect to constant-time implementations? Firstly, know that Moore's law will never help you. CPUs may get faster, but the amount of data that we need to process is increasing just as fast, if not faster. So you can't assume that the slower, constant time code will become viable in time - you have to be constant time from the start.

Even having a non-constant time implementation is a danger. There are usually two sides to a protocol and I may not control the software on the other side. If I specify AES in a protocol then I have to consider that it may well be a non-constant time implementation. I just made up the term ‘implementation ecosystem’ for this. AES is very difficult to implement securely in software: good implementations are still still topics for research papers. So the implementation ecosystem for AES is terrible! There are lots and lots of side-channel vulnerable implementations out there because no normal person, given the AES spec, will produce a secure implementation.

If we're aiming for a 128-bit security level then that possibility is a much larger problem than many other, more traditional crypto concerns.

So, for new primitives, you may want to produce solid implementations for different platforms to seed the implementation ecosystem. Not just reference implementations, but ones that are good enough that they dominate the set of implementations. For example, if I specify curve25519 in a protocol, I can be pretty sure that everyone is going to be using djb's reference code. That's a major advantage.

You should consider what an implementation is going to look like when designing. Of course, building a solid implementation will make sure that you end up considering this, so that's another good reason. There are certain patterns that are inherently dangerous. Square-and-multiply loops for example. You should recognise that and make sure that even the description of the algorithm includes counter measures. Binary fields are another which are very likely to result in non-constant time code.

Lastly, please don't assume that CPU changes are going to solve your constant-time or performance problems. Intel have added specific instructions for AES and binary fields in their latest chips and, while that does have some benefit, they will be a small fraction of all chips for a very long time. The chance of both sides of a connection having these chips is even smaller.

Cryptographic Room 101.

Room 101 is a fairly long running British TV show where people nominate things they dislike to put into Room 101, which banishes them from the world. Based on the above I've a number of things that I'd to put into Room 101, from least controversial to most. These are more specific points than the general ones that I just made and I'd like to see if anyone in the room disagrees with me!

1. MAC then Encrypt.

I hope that everyone agrees on this. Of course, it's pretty ubiquitous in the world today. Just because I want to banish something doesn't mean that I'm not going to spend years dealing with it in the future!

2. CBC mode.

With all the problems of chosen plaintexts, padding oracles etc I think it's time to get rid of CBC mode forever. Although it's possible to implement it securely, it's been done wrong for so long and is so easy to mess up, I think it's best to get rid of it.

3. ‘Sudden death’ entropy failure: plain DSA.

DSA (and ECDSA) has a very unfortunate property that an entropy failure leaks the private key. Even a slight bias over many signatures can be exploited. This is ridiculous. As Nadia and Debian have demonstrated, entropy failures do happen and they are inherently very bad. But that's not a reason to amplify them! By hashing in the private key and message, this problem can be avoided. So please consider what happens when your nonces are actually ntwices.

4. AES-GCM.

I've saved the really controversial one for last!

AES-GCM so easily leads to timing side-channels that I'd like to put it into Room 101. It took a decade of research to produce solid, high-speed, constant time AES implementations and they are very complex. In that decade, many, many non-constant time AES implementations have found their way into everything, poisoning the ecosystem when it comes to using AES.

I haven't seen any research on extracting the key from GHASH, but I've certainly seen vulnerable implementations and there's every reason to believe that it's possible. Most GHASH implementations look like AES implementations from 10 years ago. I'm aware of one, reasonable, constant-time AES-GCM implementation (Käsper and Schwabe, CHES 2009), but it runs at 22 cycles/byte on a Core2.

If you have a recent Intel chip, and software that implements AES-GCM using the specific instructions provided, then it's great. But most chips don't have that and I think it would have been much more preferable to pick an AEAD that's easy to securely implement everywhere and then speed it up on certain chips.

But it's still much better than AES-CBC I guess!

(At this point I cycled back to talking about the larger transport security issues and how browsers are addressing them with HSTS, mixed script blocking etc. But I've written about that before.)