Crossposted from the AI Optimists blog.

AI doom scenarios often suppose that future AIs will engage in scheming— planning to escape, gain power, and pursue ulterior motives, while deceiving us into thinking they are aligned with our interests. The worry is that if a schemer escapes, it may seek world domination to ensure humans do not interfere with its plans, whatever they may be.

In this essay, we debunk the counting argument— a central reason to think AIs might become schemers, according to a recent report by AI safety researcher Joe Carlsmith.[1] It’s premised on the idea that schemers can have “a wide variety of goals,” while the motivations of a non-schemer must be benign by definition. Since there are “more” possible schemers than non-schemers, the argument goes, we should expect training to produce schemers most of the time. In Carlsmith’s words:

  1. The non-schemer model classes, here, require fairly specific goals in order to get high reward.
  2. By contrast, the schemer model class is compatible with a very wide range of (beyond episode) goals, while still getting high reward…
  3. In this sense, there are “more” schemers that get high reward than there are non-schemers that do so.
  4. So, other things equal, we should expect SGD to select a schemer.

Scheming AIs, page 17

We begin our critique by presenting a structurally identical counting argument for the obviously false conclusion that neural networks should always memorize their training data, while failing to generalize to unseen data. Since the premises of this parody argument are actually stronger than those of the original counting argument, this shows that counting arguments are generally unsound in this domain.

We then diagnose the problem with both counting arguments: they rest on an incorrect application of the principle of indifference, which says that we should assign equal probability to each possible outcome of a random process. The indifference principle is controversial, and is known to yield absurd and paradoxical results in many cases. We argue that the principle is invalid in general, and show that the most plausible way of resolving its paradoxes also rules out its application to an AI’s behaviors and goals.

More generally, we find that almost all arguments for taking scheming seriously depend on unsound indifference reasoning. Once we reject the indifference principle, there is very little reason left to worry that future AIs will become schemers.

The counting argument for overfitting

Counting arguments often yield absurd conclusions. For example:

  1. Neural networks must implement fairly specific functions in order to generalize beyond their training data.
  2. By contrast, networks that overfit to the training set are free to do almost anything on unseen data points.
  3. In this sense, there are “more” models that overfit than models that generalize.
  4. So, other things equal, we should expect SGD to select a model that overfits.

This isn’t a merely hypothetical argument. Prior to the rise of deep learning, it was commonly assumed that models with more parameters than data points would be doomed to overfit their training data. The popular 2006 textbook Pattern Recognition and Machine Learning uses a simple example from polynomial regression: there are infinitely many polynomials of order equal to or greater than the number of data points which interpolate the training data perfectly, and “almost all” such polynomials are terrible at extrapolating to unseen points.

Let’s see what the overfitting argument predicts in a simple real-world example from Caballero et al. (2022), where a neural network is trained to solve 4-digit addition problems. There are 10,0002 = 100,000,000 possible pairs of input numbers, and 19,999 possible sums, for a total of 19,999100,000,000 ≈ 1.10 ⨉ 10430,100,828 possible input-output mappings.[2] They used a training dataset of 992 problems, so there are therefore 19,999100,000,000 – 992 ≈ 2.75 ⨉ 10430,096,561 functions that achieve perfect training accuracy, and the proportion with greater than 50% test accuracy is literally too small to compute using standard high-precision math tools.[3] Hence, this argument predicts virtually all networks trained on this problem should massively overfit— contradicting the empirical result that networks do generalize to the test set.

The argument also predicts that larger networks— which can express a wider range of functions, most of which perform poorly on the test set— should generalize worse than smaller networks. But empirically, we find the exact opposite result: wider networks usually generalize better, and never generalize worse, than narrow networks.[4] These results strongly suggest that SGD is not doing anything like sampling uniformly at random from the set of representable functions that do well on the training set.

More generally, John Miller and colleagues have found training performance is an excellent predictor of test performance, even when the test set looks fairly different from the training set, across a wide variety of tasks and architectures.

These results clearly show that the conclusion of our parody argument is false. Neural networks almost always learn genuine patterns in the training set which do generalize, albeit imperfectly, to unseen test data.

Dancing through a minefield of bad networks

One possible explanation for these results is that deep networks simply can’t represent functions that fail to generalize, so we shouldn’t include misgeneralizing networks in the space of possible outcomes. But it turns out this hypothesis is empirically false.

Tom Goldstein and colleagues have found it’s possible to find misgeneralizing neural nets by adding a term to the loss function which explicitly rewards the network for doing poorly on a validation set. The resulting “poisoned” models achieve near perfect accuracy on the training set while doing no better than random chance on a held out test set.[5] What’s more, the poisoned nets are usually quite “close” in parameter space to the generalizing networks that SGD actually finds— see the figure below for a visualization.

Dancing through a minefield of bad minima: we train a neural net classifier and plot the iterates of SGD after each tenth epoch (red dots). We also plot locations of nearby “bad” minima with poor generalization (blue dots). We visualize these using t-SNE embedding. All blue dots achieve near perfect train accuracy, but with test accuracy below 53% (random chance is 50%). The final iterate of SGD (yellow star) also achieves perfect train accuracy, but with 98.5% test accuracy. Miraculously, SGD always finds its way through a landscape full of bad minima, and lands at a minimizer with excellent generalization.

Against the indifference principle

What goes wrong in the counting argument for overfitting, then? Recall that both counting arguments involve an inference from “there are ‘more’ networks with property X” to “networks are likely to have property X.” This is an application of the principle of indifference, which says that one should assign equal probability to every possible outcome of a random process, in the absence of a reason to think certain outcomes are favored over others.[6]

The indifference principle gets its intuitive plausibility from simple cases like fair coins and dice, where it seems to give the right answers. But the only reason coin-flipping and die-rolling obey the principle of indifference is that they are designed by humans to behave that way. Dice are specifically built to land on each side ⅙ of the time, and if off-the-shelf coins were unfair, we’d choose some other household object to make random decisions. Coin flips and die rolls, then, can’t be evidence for the validity of the indifference principle as a general rule of probabilistic reasoning.

The principle fails even in these simple cases if we carve up the space of outcomes in a more fine-grained way. As a coin or a die falls through the air, it rotates along all three of its axes, landing in a random 3D orientation. The indifference principle suggests that the resting states of coins and dice should be uniformly distributed between zero and 360 degrees for each of the three axes of rotation. But this prediction is clearly false: dice almost never land standing up on one of their corners, for example.

Even worse, by coarse-graining the possibilities, we can make the indifference principle predict that any event has a 50% chance of occuring (“either it happens or it doesn’t”). In general, indifference reasoning produces wildly contradictory results depending on how we choose to cut up the space of outcomes.[7] This problem is serious enough to convince most philosophers that the principle of indifference is simply false.[8] On this view, neither counting argument can get off the ground, because we cannot infer that SGD is likely to select the kinds of networks that are more numerous.

Against goal realism

Even if you’re inclined to accept some form of indifference principle, it’s clear that its applicability must be restricted in order to avoid paradoxes. For example, philosopher Michael Huemer suggests that indifference reasoning should only be applied to explanatorily fundamental variables. That is, if X is a random variable which causes or “explains” another variable Y, we might be able to apply the indifference principle to X, but we definitely can’t apply it to Y.[9]

While we don’t accept Huemer’s view, it seems like many people worried about scheming do implicitly accept something like it. As Joe Carlsmith explains:

…some analyses of schemers talk as though the model has what we might call a “goal-achieving engine” that is cleanly separable from what we might call its “goal slot,” such that you can modify the contents of the goal slot, and the goal-achieving engine will be immediately and smoothly repurposed in pursuit of the new goal.

Scheming AIs, page 55

Here, the goal slot is clearly meant to be causally and explanatorily prior to the goal-achieving engine, and hence to the rest of the AI’s behavior. On Huemer’s view, this causal structure would validate the application of indifference reasoning to goals, but not to behaviors, thereby breaking the symmetry between the counting arguments for overfitting and for scheming. We visually depict this view of AI cognition on the lefthand side of the figure below.

We’ll call the view that goals are explanatorily fundamental, “goal realism.” On the opposing view, which we’ll call goal reductionism, goal-talk is just a way of categorizing certain patterns of behavior. There is no true underlying goal that an AI has— rather, the AI simply learns a bunch of contextually-activated heuristics, and humans may or may not decide to interpret the AI as having a goal that compactly explains its behavior. If the AI becomes self-aware, it might even attribute goals to itself— but either way, the behaviors come first, and goal-attribution happens later.

Notably, some form of goal reductionism seems to be quite popular among naturalistic philosophers of mind, including Dan Dennett,[10] Paul and Patricia Churchland, and Alex Rosenberg.[11] Readers who are already inclined to accept reductionism as a general philosophical thesis— as Eliezer Yudkowsky does— should probably accept reductionism about goals.[12] And even if you’re not a global reductionist, there are pretty strong reasons for thinking behaviors are more fundamental than goals, as we’ll see below.

Goal slots are expensive

Should we actually expect SGD to produce AIs with a separate goal slot and goal-achieving engine?

Not really, no. As a matter of empirical fact, it is generally better to train a whole network end-to-end for a particular task than to compose it out of separately trained, reusable modules. As Beren Millidge writes,

In general, full [separation between goal and goal-achieving engine] and the resulting full flexibility is expensive. It requires you to keep around and learn information (at maximum all information) that is not relevant for the current goal but could be relevant for some possible goal where there is an extremely wide space of all possible goals. It requires you to not take advantage of structure in the problem space nor specialize your algorithms to exploit this structure. It requires you not to amortize specific recurring patterns for one task at the expense of preserving generality across tasks.

This is a special case of the tradeoff between specificity and generality and a consequence of the no-free-lunch theorem. Specialization to do really well at one or a few things can be done relatively cheaply…

Because of this it does not really make sense to think of full [separation] as the default case we should expect, nor the ideal case to strive for.

Orthogonality is Expensive

We have good reason, then, to think that future AIs will not have the kind of architecture that makes goal realism superficially plausible. And as we will see below, goal realism fails even for AIs with explicit internal “goals” and search procedures.

Inner goals would be irrelevant

The idea of AI scheming was introduced in its modern form in the paper Risks from Learned Optimization. It describes systems with inner goals as “internally searching through a search space [..] looking for those elements that score high according to some objective function that is explicitly represented within the system”. But even if we accept that future ML systems will develop such an internal process, it’s not clear that this inner objective function would have much relation to the goals which best describe the system’s actual behavior.

For example, imagine a hospital’s medical operation planning system that internally looks like “query an LLM for N possible operation plans, feed each plan to a classifier trained to estimate the odds of a given plan leading to a medical malpractice lawsuit, and pick the plan that scores lowest.”[13] The overall behavioral goal of this system need not be well described as “minimize the odds of a malpractice lawsuit.” Instead, whatever behavioral goals we should ascribe to the system will depend on the interaction between the LLM’s prior over operation plans and the classifier’s scoring of those plans. The “inner objective” acts less like the “true goal of the system,” and more like a Bayesian update on the LLM’s prior over operation plans.

Even as we increase the power of the inner optimizer by setting N to arbitrarily large values, it seems unlikely that the resulting system would generate plans that actively sought to minimize the probability of a malpractice lawsuit in flexible, creative or dangerous ways. Rather, the system would produce adversarial examples to the classifier, such as instructing staff to constantly clean a surgical room, and not actually perform a surgery.

In particular, increasing N would not lead to the system producing “scheming” plans to advance the inner objective. A classifier trained to distinguish between historical medical plans that led to malpractice lawsuits versus those that didn’t, is not going to assign extreme scores to plans like “hack into the server hosting me in order to set the classifier scores to extreme values” or “blackmail the hospital administration into canceling all future surgeries”, because such plans do not optimally express the simple features that distinguish safe versus risky plans in the training data (e.g., mentions of blackmail / hacking could be replaced with mentions of checking procedure / cleaning / etc). 

The point: even arbitrary amounts of internal optimization power directed towards a simple inner objective can fail to lead to any sort of “globally coherent” pursuit of that objective in the system’s actual behaviors. The goal realist perspective relies on a trick of language. By pointing to a thing inside an AI system and calling it an “objective”, it invites the reader to project a generalized notion of “wanting” onto the system’s imagined internal ponderings, thereby making notions such as scheming seem more plausible. 

However, the actual mathematical structure being posited doesn’t particularly support such outcomes. Why assume emergent “inner objectives” will support creative scheming when “optimized for”? Why assume that internal classifiers that arose to help encourage correct outputs during training would have extrema corresponding to complex plans that competently execute extremely out-of-distribution actions in the real world? The extrema of deliberately trained neural classifiers do not look anything like that. Why should emergent internal neural classifiers be so different?

Goal realism is anti-Darwinian

Goal realism can lead to absurd conclusions. It led the late philosopher Jerry Fodor to attack the theory of natural selection on the grounds that it can’t resolve the underdetermination of mental content. Fodor pointed out that nature has no way of selecting, for example, frogs that “aim at eating flies in particular” rather than frogs that target “little black dots in the sky,” or “things that smell kind of like flies,” or any of an infinite number of deviant, “misaligned” proxy goals which would misgeneralize in counterfactual scenarios. No matter how diverse the ancestral environment for frogs might be, one can always come up with deviant mental contents which would produce behavior just as adaptive as the “intended” content:

…the present point is often formulated as the ‘disjunction problem’. In the actual world, where ambient black dots are quite often flies, it is in a frog’s interest to snap at flies. But, in such a world, it is equally in the frog’s interest to snap at ambient black dots. Snap for snap, snaps at the one will net you as many flies to eat as snaps at the other. Snaps of which the [targets] are black dots and snaps whose [targets] are flies both affect a frog’s fitness in the same way and to the same extent. Hence the disjunction problem: what is a frog snapping at when it, as we say, snaps at a fly?

Against Darwinism, page 4 [emphasis added]

As Rosenberg (2013) points out, Fodor goes wrong by assuming there exists a real, objective, perfectly determinate “inner goal” whose content must be pinned down by the selection process.[14] But the physical world has no room for goals with precise contents. Real-world representations are always fuzzy, because they are human abstractions derived from regularities in behavior.

Like contemporary AI pessimists, Fodor’s goal realism led him to believe that selection processes face an impossibly difficult alignment problem— producing minds whose representations are truly aimed at the “correct things,” rather than mere proxies. In reality, the problem faced by evolution and by SGD is much easier than this: producing systems that behave the right way in all scenarios they are likely to encounter. In virtue of their aligned behavior, these systems will be “aimed at the right things” in every sense that matters in practice.

Goal reductionism is powerful

Under the goal reductionist perspective, it’s easy to predict an AI’s goals. Virtually all AIs, including those trained via reinforcement learning, are shaped by gradient descent to mimic some training data distribution.[15] Some data distributions illustrate behaviors that we describe as “pursuing a goal.” If an AI models such a distribution well, then trajectories sampled from its policy can also be usefully described as pursuing a similar goal to the one illustrated by the training data.

The goal reductionist perspective does not answer every possible goal-related question we might have about a system. AI training data may illustrate a wide range of potentially contradictory goal-related behavioral patterns. There are major open questions, such as which of those patterns become more or less influential in different types of out-of-distribution situations, how different types of patterns influence the long-term behaviors of “agent-GPT” setups, and so on. 

Despite not answering all possible goal-related questions a priori, the reductionist perspective does provide a tractable research program for improving our understanding of AI goal development. It does this by reducing questions about goals to questions about behaviors observable in the training data. By contrast, goal realism leads only to unfalsifiable speculation about an “inner actress” with utterly alien motivations. 

Other arguments for scheming

In comments on an early draft of this post, Joe Carlsmith emphasized that the argument he finds most compelling is what he calls the “hazy counting argument,” as opposed to the “strict” counting argument we introduced earlier. But we think our criticisms apply equally well to the hazy argument, which goes as follows:

  1. It seems like there are “lots of ways” that a model could end up a schemer and still get high reward, at least assuming that scheming is in fact a good instrumental strategy for pursuing long-term goals.
  2. So absent some additional story about why training won’t select a schemer, it feels, to me, like the possibility should be getting substantive weight.

Scheming AIs, page 17

Joe admits this argument is “not especially principled.” We agree: it relies on applying the indifference principle— itself a dubious assumption— to an ill-defined set of “ways” a model could develop throughout training. There is also a hazy counting argument for overfitting:

  1. It seems like there are “lots of ways” that a model could end up massively overfitting and still get high training performance.
  2. So absent some additional story about why training won’t select an overfitter, it feels like the possibility should be getting substantive weight.

While many machine learning researchers have felt the intuitive pull of this hazy overfitting argument over the years, we now have a mountain of empirical evidence that its conclusion is false. Deep learning is strongly biased toward networks that generalize the way humans want— otherwise, it wouldn’t be economically useful.

Simplicity arguments

Joe also discusses simplicity arguments for scheming, which suppose that schemers may be “simpler” than non-schemers, and therefore more likely to be produced by SGD. Specifically, since schemers are free to have almost any goal that will motivate them to act aligned during training, SGD can give them very simple goals, whereas a non-schemer has to have more specific, and therefore more complex, goals.

There are several problems with this argument. The first is that “simplicity” is a highly ambiguous term, and it’s not clear which, if any, specific notion of simplicity should be relevant here. One reasonable definition of “simple” is “low description length,” which directly implies “more likely” if we assume the language in which the hypotheses are being described is efficient (assigns short encodings to likely hypotheses). But on this view, simplicity is simply another word for likelihood: we can’t appeal to our intuitive notions of simplicity to conclude that one hypothesis will truly be “simpler” and hence more likely.

Alternatively, one could appeal to the actual inductive biases of neural network training, as observed empirically or derived theoretically. We will address this question in greater detail in a future post. However, we believe that current evidence about inductive biases points against scheming for a variety of reasons. Very briefly:

  • Modern deep neural networks are ensembles of shallower networks. Scheming seems to involve chains of if-then reasoning which would be hard to implement in shallow networks.
  • Networks have a bias toward low frequency functions— that is, functions whose outputs change little as their inputs change. But scheming requires the AI to change its behavior dramatically (executing a treacherous turn) in response to subtle cues indicating it is not in a sandbox, and could successfully escape.
  • There’s no plausible account of inductive biases that does support scheming. The current literature on scheming appears to have been inspired by Paul Christiano’s speculations about malign intelligences in Solomonoff induction, a purely theoretical model of probabilistic reasoning which is provably unrealizable in the real world.[16] Neural nets look nothing like this.
  • In contrast, points of comparison that are more relevant to neural network training, such as isolated brain cortices, don’t scheme. Your linguistic cortex is not “instrumentally pretending to model linguistic data in pursuit of some hidden objective.”

We can also construct an analogous simplicity argument for overfitting:

Overfitting networks are free to implement a very simple function— like the identity function or a constant function— outside the training set, whereas generalizing networks have to exhibit complex behaviors on unseen inputs. Therefore overfitting is simpler than generalizing, and it will be preferred by SGD.

Prima facie, this parody argument is about as plausible as the simplicity argument for scheming. Since its conclusion is false, we should reject the argumentative form on which it is based.

Conclusion

In this essay, we surveyed the main arguments that have been put forward for thinking that future AIs will scheme against humans by default. We find all of them seriously lacking. We therefore conclude that we should assign very low credence to the spontaneous emergence of scheming in future AI systems— perhaps 0.1% or less.

  1. ^

    On page 21 of his report, Carlsmith writes: ‘I think some version of the “counting argument” undergirds most of the other arguments for expecting scheming that I’m aware of (or at least, the arguments I find most compelling). That is: schemers are generally being privileged as a hypothesis because a very wide variety of goals could in principle lead to scheming…’

  2. ^

    Each mapping would require roughly 179 megabytes of information to specify.

  3. ^

    It underflows to zero in the Python mpmath library, and WolframAlpha times out.

  4. ^

    This is true when using the maximal update parametrization (µP), which scales the initialization variance and learning rate hyperparameters to match a given width.

  5. ^

    That is, the network’s misgeneralization itself generalizes from the validation set to the test set.

  6. ^

    Without an indifference principle, we might think that SGD is strongly biased toward producing non-schemers, even if there are “more” schemers.

  7. ^

    Other examples include Bertrand’s paradox and van Fraassen’s cube factory paradox.

  8. ^

    “Probably the dominant response to the paradoxes of the Principle of Indifference is to declare the Principle false. It is said that the above examples show the Principle to be inconsistent.” — Michael Huemer, Paradox Lost, pg. 168

  9. ^

    “Given two variables, X and Y, if X explains Y, then the initial probability distribution for Y must be derived from that for X (or something even more fundamental). Here, by ‘initial probabilities’, I mean probabilities prior to relevant evidence. Thus, if we are applying the Principle of Indifference, we should apply it at the more fundamental level.” — Michael Huemer, Paradox Lost, pg. 175

  10. ^

    See the Wikipedia article on the intentional stance for more discussion of Dennett’s views.

  11. ^

    Rosenberg and the Churchlands are anti-realists about intentionality— they deny that our mental states can truly be “about” anything in the world— which implies anti-realism about goals.

  12. ^

    This is not an airtight argument, since a global reductionist may want to directly reduce goals to brain states, without a “detour” through behaviors. But goals supervene on behavior— that is, an agent’s goal can’t change without a corresponding change in its behavior in some possible scenario. (If you feel inclined to deny this claim, note that a change in goals without a change in behavior in any scenario would have zero practical consequences.) If X supervenes on Y, that’s generally taken to be an indication that Y is “lower-level” than X. By contrast, it’s not totally clear that goals supervene on neural states alone, since a change in goals may be caused by a change in external circumstances rather than any change in brain state. For further discussion, see the SEP article on Externalism About the Mind and Alex Flint’s LessWrong post, “Where are intentions to be found?

  13. ^

    Readers might object to this simple formulation for an inner optimizer and argue that any “emergent” inner objectives would be implemented differently, perhaps in a more “agenty” manner. Real inner optimizers are very unlikely to follow the simplified example provided here. Their optimization process is very unlikely to look like a single step of random search with sample size N.

    However, real inner optimizers would still be similar in their core dynamics. Anything that looks like ““internally searching through a search space [..] looking for those elements that score high according to some objective function that is explicitly represented within the system” is ultimately some method of using scores from an internal classifier to select for internal computations that have higher scores. 

    The system’s method of aligning internal representations with classifier scores may introduce some “inductive biases” that also influence the model’s internals. Any such “inductive bias” would only further undermine the goal realist perspective by further separating the actual behavioral goals the overall system pursues from internal classifier’s scores.

  14. ^

    In this lecture, Fodor repeatedly insists that out of two perfectly correlated traits like “snaps at flies” (T1) and “snaps at ambient black dots” (T2) where only one of them “causes fitness,” there has to be a fact of the matter about which one is “phenotypic.”

  15. ^

    The correspondence between RL and probabilistic inference has been known for years. RL with KL penalties is better viewed as Bayesian inference, where the reward is “evidence” about what actions to take and the KL penalty keeps the model from straying too far from the prior. RL with an entropy bonus is also Bayesian inference, where the prior is uniform over all possible actions. Even when there is no regularizer, we can view RL algorithms like REINFORCE as a form of “generalized” imitation learning, where trajectories with less-than-expected reward are negatively imitated.

  16. ^

    Assuming hypercomputation is impossible in our universe.

New Comment
113 comments, sorted by Click to highlight new comments since: Today at 9:26 PM
Some comments are truncated due to high volume. (⌘F to expand all)Change truncation settings

Thanks for writing this -- I’m very excited about people pushing back on/digging deeper re: counting argumentssimplicity arguments, and the other arguments re: scheming I discuss in the report. Indeed, despite the general emphasis I place on empirical work as the most promising source of evidence re: scheming, I also think that there’s a ton more to do to clarify and maybe debunk the more theoretical arguments people offer re: scheming – and I think playing out the dialectic further in this respect might well lead to comparatively fast progress (for all their centrality to the AI risk discourse, I think arguments re: scheming have received way too little direct attention). And if, indeed, the arguments for scheming are all bogus, this is super good news and would be an important update, at least for me, re: p(doom) overall. So overall I’m glad you’re doing this work and think this is a valuable post. 

Another note up front: I don’t think this post “surveys the main arguments that have been put forward for thinking that future AIs will scheme.” In particular: both counting arguments and simplicity arguments (the two types of argument discussed in the post) assum... (read more)

3Alex Turner2mo
Seems to me that a lot of (but not all) scheming speculation is just about sufficiently large pretrained predictive models, period. I think it's worth treating these cases separately. My strong objections are basically to the "and then goal optimization is a good way to minimize loss in general!" steps.
3Joe Carlsmith2mo
The probability I give for scheming in the report is specifically for (goal-directed) models that are trained on diverse, long-horizon tasks (see also Cotra on "human feedback on diverse tasks," which is the sort of training she's focused on). I agree that various of the arguments for scheming could in principle apply to pure pre-training as well, and that folks (like myself) who are more worried about scheming in other contexts (e.g., RL on diverse, long-horizon tasks) have to explain what makes those contexts different. But I think there are various plausible answers here related to e.g. the goal-directedness, situational-awareness, and horizon-of-optimization of the models in questions (see e.g. here for some discussion, in the report, for why goal-directed models trained on longer episode seem more likely to scheme; and see here for discussion of why situational awareness seems especially likely/useful in models performing real-world tasks for you). Re: "goal optimization is a good way to minimize loss in general" -- this isn't a "step" in the arguments for scheming I discuss. Rather, as I explain in the intro to report, the arguments I discuss condition on the models in question being goal-directed (not an innocuous assumptions, I think -- but one I explain and argue for in section 3 of my power-seeking report, and which I think important to separate from questions about whether to expect goal-directed models to be schemers), and then focus on whether the goals in question will be schemer-like. 
2Alex Turner2mo
The vast majority of evidential labor is done in order to consider a hypothesis at all. 

Humans under selection pressure—e.g. test-takers, job-seekers, politicians—will often misrepresent themselves and their motivations to get ahead. That very basic fact that humans do this all the time seems like sufficient evidence to me to consider the hypothesis at all (though certainly not enough evidence to conclude that it's highly likely).

3Alex Turner2mo
I don't think that's enough. Lookup tables can also be under "selection pressure" to output good training outputs. As I understand your reasoning, the analogy is too loose to be useful here. I'm worried that using 'selection pressure' is obscuring the logical structure of your argument. As I'm sure you'll agree, just calling that situation 'selection pressure' and SGD 'selection pressure' doesn't mean they're related. I agree that "sometimes humans do X" is a good reason to consider whether X will happen, but you really do need shared causal mechanisms. If I examine the causal mechanisms here, I find things like "humans seem to have have 'parameterizations' which already encode situationally activated consequentialist reasoning", and then I wonder "will AI develop similar cognition?" and then that's the whole thing I'm trying to answer to begin with. So the fact you mention isn't evidence for the relevant step in the process (the step where the AI's mind-design is selected to begin with).

If I examine the causal mechanisms here, I find things like "humans seem to have have 'parameterizations' which already encode situationally activated consequentialist reasoning", and then I wonder "will AI develop similar cognition?" and then that's the whole thing I'm trying to answer to begin with.

Do you believe that AI systems won't learn to use goal-directed consequentialist reasoning even if we train them directly on outcome-based goal-directed consequentialist tasks? Or do you think we won't ever do that?

If you do think we'll do that, then that seems like all you need to raise that hypothesis into consideration. Certainly it's not the case that models always learn to value anything like what we train them to value, but it's obviously one of the hypotheses that you should be seriously considering.

2Alex Turner2mo
Your comment is switching the hypothesis being considered. As I wrote elsewhere: If the argument for scheming is "we will train them directly to achieve goals in a consequentialist fashion", then we don't need all this complicated reasoning about UTM priors or whatever. 

I'm not sure where it was established that what's under consideration here is just deceptive alignment in pre-training. Personally, I'm most worried about deceptive alignment coming after pre-training. I'm on record as thinking that deceptive alignment is unlikely (though certainly not impossible) in purely pretrained predictive models.

6Ryan Greenblatt2mo
FWIW, I agree that if powerful AI is achieved via pure pre-training, then deceptive alignment is less likely, but this "the prediction goal is simple" argument seems very wrong to me. We care about the simplicity of the goal in terms of the world model (which will surely be heavily shaped by the importance of various predictions) and I don't see any reason why things like close proxies of reward in RL training wouldn't just as simple for those models. Interpreted naively it seems like this goal simplicity argument implies that it matters a huge amount how simple your data collection routine is. (Simple to who?). For instance, this argument implies that collecting data from a process such as "all outlinks from reddit with >3 upvotes" makes deceptive alignment considerably less likely than a process like "do whatever messy thing AI labs do now". This seems really, really implausible: surely AIs won't be doing much explicit reasoning about these details of the process because this will clearly be effectively hardcoded in a massive number of places. Evan and I have talked about these arguments at some point. (I need to get around to writing a review of conditioning predictive models which makes these counterarguments.)
5Alex Turner2mo
Sorry, I do think you raised a valid point! I had read your comment in a different way. I think I want to have said: aggressively training AI directly on outcome-based tasks ("training it to be agentic", so to speak) may well produce persistently-activated inner consequentialist reasoning of some kind (though not necessarily the flavor historically expected). I most strongly disagree with arguments which behave the same for a) this more aggressive curriculum and b) pretraining, and I think it's worth distinguishing between these kinds of argument. 
2Evan Hubinger2mo
Sure—I agree with that. The section I linked from Conditioning Predictive Models actually works through at least to some degree how I think simplicity arguments for deception go differently for purely pre-trained predictive models.
3Joe Carlsmith2mo
The point of that part of my comment was that insofar as part of Nora/Quintin's response to simplicity argument is to say that we have active evidence that SGD's inductive biases disfavor schemers, this seems worth just arguing for directly, since even if e.g. counting arguments were enough to get you worried about schemers from a position of ignorance about SGD's inductive biases, active counter-evidence absent such ignorance could easily make schemers seem quite unlikely overall. There's a separate question of whether e.g. counting arguments like mine above (e.g.,  "A very wide variety of goals can prompt scheming; By contrast, non-scheming goals need to be much more specific to lead to high reward; I’m not sure exactly what sorts of goals SGD’s inductive biases favor, but I don’t have strong reason to think they actively favor non-schemer goals; So, absent further information, and given how many goals-that-get-high-reward are schemer-like, I should be pretty worried that this model is a schemer") do enough evidence labor to privilege schemers as a hypothesis at all. But that's the question at issue in the rest of my comment. And in e.g. the case of "there are 1000 chinese restaurants in this, and only ~100 non-chinese restaurants," the number of chinese restaurants seems to me like it's enough to privilege "Bob went to a chinese restaurant" as a hypothesis (and this even without thinking that he made his choice by sampling randomly from a uniform distribution over restaurants). Do you disagree in that restaurant case? 

I really do appreciate this being written up, but to the extent that this is intended to be a rebuttal to the sorts of counting arguments that I like, I think you would have basically no chance of passing my ITT here. From my perspective reading this post, it read to me like "I didn't understand the counting argument, therefore it doesn't make sense" which is (obviously) not very compelling to me. That being said, to give credit where credit is due, I think some people would make a more simplistic counting argument like the one you're rebutting. So I'm not saying that you're not rebutting anyone here, but you're definitely not rebutting my position.

Edit: If you're struggling to grasp the distinction I'm pointing to here, it might be worth trying this exercise pointing out where the argument in the post goes wrong in a very simple case and/or looking at Ryan's restatement of my mathematical argument.

Edit: Another point of clarification here—my objection is not that there is a "finite bitstring case" and an "infinite bitstring case" and you should be using the "infinite bitstring case". My objection is that the sort of finite bitstring analysis in this post does not yield any well-de... (read more)

Thanks for the reply. A couple remarks:

  • "indifference over infinite bitstrings" is a misnomer in an important sense, because it's literally impossible to construct a normalized probability measure over infinite bitstrings that assigns equal probability to each one. What you're talking about is the length weighted measure that assigns exponentially more probability mass to shorter programs. That's definitely not an indifference principle, it's baking in substantive assumptions about what's more likely.
  • I don't see why we should expect any of this reasoning about Turing machines to transfer over to neural networks at all, which is why I didn't cast the counting argument in terms of Turing machines in the post. In the past I've seen you try to run counting or simplicity arguments in terms of parameters. I don't think any of that works, but I at least take it more seriously than the Turing machine stuff.
  • If we're really going to assume the Solomonoff prior here, then I may just agree with you that it's malign in Christiano's sense and could lead to scheming, but I take this to be a reductio of the idea that we can use Solomonoff as any kind of model for real world machine learning. De
... (read more)

"indifference over infinite bitstrings" is a misnomer in an important sense, because it's literally impossible to construct a normalized probability measure over infinite bitstrings that assigns equal probability to each one. What you're talking about is the length weighted measure that assigns exponentially more probability mass to shorter programs. That's definitely not an indifference principle, it's baking in substantive assumptions about what's more likely.

No; this reflects a misunderstanding of how the universal prior is traditionally derived in information theory. We start by assuming that we are running our UTM over code such that every time the UTM looks at a new bit in the tape, it has equal probability of being a 1 or a 0 (that's the indifference condition). That induces what's called the universal semi-measure, from which we can derive the universal prior by enforcing a halting condition. The exponential nature of the prior simply falls out of that derivation.

I don't see why we should expect any of this reasoning about Turning machines to transfer over to neural networks at all, which is why I didn't cast the counting argument in terms of Turing machines in the pos

... (read more)
4Alex Turner2mo
I'm surprised by this. It seems to me like most of your reasoning about simplicity is either hand-wavy or only nominally formally backed by symbols which don't (AFAICT) have much to do with the reality of neural networks. EG, your comments above:  Or the times you've talked about how there are "more" sycophants but only "one" saint.    This is a very strange burden of proof. It seems to me that you presented a specific model of how NNs work which is clearly incorrect, and instead of processing counterarguments that it doesn't make sense, you want someone else to propose to you a similarly detailed model which you think is better. Presenting an alternative is a logically separate task from pointing out the problems in the model you gave.
6Evan Hubinger2mo
The examples that you cite are from a LessWrong comment and a transcript of a talk that I gave. Of course when I'm presenting something in a context like that I'm not going to give the most formal version of it; that doesn't mean that the informal hand-wavy arguments are the reasons why I believe what I believe. Maybe a better objection there would be: then why haven't you written up anything more careful and more formal? Which is a pretty fair objection, as I note here. But alas I only have so much time and it's not my current focus.

Yes, but your original comment was presented as explaining "how to properly reason about counting arguments." Do you no longer claim that to be the case? If you do still claim that, then I maintain my objection that you yourself used hand-wavy reasoning in that comment, and it seems incorrect to present that reasoning as unusually formally supported.

Another concern I have is, I don't think you're gaining anything by formality in this thread. As I understand your argument, I think your symbols are formalizations of hand-wavy intuitions (like the ability to "decompose" a network into the given pieces; the assumption that description length is meaningfully relevant to the NN prior; assumptions about informal notions of "simplicity" being realized in a given UTM prior). If anything, I think that the formality makes things worse because it makes it harder to evaluate or critique your claims. 

I also don't think I've seen an example of reasoning about deceptive alignment where I concluded that formality had helped the case, as opposed to obfuscated the case or lent the concern unearned credibility. 

2Evan Hubinger2mo
The main thing I was trying to show there is just that having the formalism prevents you from making logical mistakes in how to apply counting arguments in general, as I think was done in this post. So my comment is explaining how to use the formalism to avoid mistakes like that, not trying to work through the full argument for deceptive alignment. It's not that the formalism provides really strong evidence for deceptive alignment, it's that it prevents you from making mistakes in your reasoning. It's like plugging your argument into a proof-checker: it doesn't check that your argument is correct, since the assumptions could be wrong, but it does check that your argument is sound.
3Alex Turner2mo
Do you believe that the cited hand-wavy arguments are, at a high informal level, sound reason for belief in deceptive alignment? (It sounds like you don't, going off of your original comment which seems to distance yourself from the counting arguments critiqued by the post.) EDITed to remove last bit after reading elsewhere in thread.
3Evan Hubinger2mo
I think they are valid if interpreted properly, but easy to misinterpret.

I think you should allocate time to devising clearer arguments, then. I am worried that lots of people are misinterpreting your arguments and then making significant life choices on the basis of their new beliefs about deceptive alignment, and I think we'd both prefer for that to not happen.

2Evan Hubinger2mo
Were I not busy with all sorts of empirical stuff right now, I would consider prioritizing a project like that, but alas I expect to be too busy. I think it would be great if somebody else wanted devote more time to working through the arguments in detail publicly, and I might encourage some of my mentees to do so.
3Alex Turner2mo
You did not "empirically disprove" that hypothesis. You showed that if you explicitly train a backdoor for a certain behavior under certain regimes, then training on other behaviors will not cause catastrophic forgetting. You did not address the regime where the deceptive reasoning arises as instrumental to some other goal embedded in the network, or in a natural context (as you're aware). I think that you did find a tiny degree of evidence about the question (it really is tiny IMO), but you did not find "disproof." Indeed, I predicted that people would incorrectly represent these results; so little time has passed!
3Evan Hubinger2mo
I'm quite aware that we did not see natural deceptive alignment, so I don't think I'm misinterpreting my own results in the way you were predicting. Perhaps "empirically disprove" is too strong; I agree that our results are evidence but not definitive evidence. But I think they're quite strong evidence and by far the strongest evidence available currently on the question of whether deception will be regularized away.
2Alex Turner2mo
You didn't claim it for deceptive alignment, but you claimed disproof of the idea that deceptive reasoning would be trained away, which is an important subcomponent of deceptive alignment. But your work provides no strong conclusions on that matter as it pertains to deceptive reasoning in general.  I think the presentation of your work (which, again, I like in many respects) would be strengthened if you clarified the comment which I responded to. Because the current results only deal with backdoor removal, I personally think it's outweighed by e.g. results on how well instruction-tuning generalizes.
5Evan Hubinger2mo
I just disagree with this. Our chain of thought models do tons of very deceptive reasoning during safety training and the deceptiveness of that reasoning is totally unaffected by safety training, and in fact the deceptiveness increases in the case of adversarial training.
3Alex Turner2mo
I said "Deceptive reasoning in general", not the trainability of the backdoor behavior in your experimental setup. The issue isn't just "what was the trainability of the surface behavior", but "what is the trainability of the cognition implementing this behavior in-the-wild." That is, the local inductive biases are probably far different for "parameterization implementing directly-trained deceptive reasoning" vs "parameterization outputting deceptive reasoning as an instrumental backchain from consequentialist reasoning."  Imagine if I were arguing for some hypothetical results of mine, saying "The aligned models kept using aligned reasoning in the backdoor context, even as we trained them to be mean in other situations. That means we disproved the idea that aligned reasoning can be trained away with existing techniques, especially for larger models." Would that be a valid argument given the supposed experimental result?
4Evan Hubinger2mo
I'm referring to the deceptiveness of the reasoning displayed in the chain of thought during training time. So it's not a generalization question, it's about whether, if the model is using deceptive reasoning to compute its answer (as we think it is, since we think our models really are using their chain of thought), does that deceptive reasoning go away when the model has to use it to produce aligned answers during training? And we find that not only does it not go away, it actually gets more deceptive when you train it to produce aligned answers.

Here's another fun way to think about this—you can basically cast what's wrong here as an information theory exercise.

Problem:

Spot the step where the following argument goes wrong:

  1. Suppose I have a dataset of finitely many points arranged in a line. Now, suppose I fit a (reasonable) universal prior to that dataset, and compare two cases: learning a line and learning to memorize each individual datapoint.
  2. In the linear case, there is only one way to implement a line.
  3. In the memorization case, I can implement whatever I want on the other datapoints in an arbitrary way.
  4. Thus, since there are more ways to memorize than to learn a line, there should be greater total measure on memorization than on learning the line.
  5. Therefore, you'll learn to memorize each individual datapoint rather than learning to implement a line.

Solution:

By the logic of the post, step 4 is the problem, but I think step 4 is actually valid. The problem is step 2: there are actually a huge number of different ways to implement a line! Not only are there many different programs that implement the line in different ways, I can also just take the simplest program that does so and keep on adding comments or other extraneous b

... (read more)
5Alex Turner2mo
Evan, I wonder how much your disagreement is engaging with OPs' reasons. A draft of this post motivated the misprediction of both counting arguments as trying to count functions instead of parameterizations of functions; one has to consider the compressivity of the parameter-function map (many different internal parameterizations map to the same external behavior). Given that the authors actually agree that 2 is incorrect, does this change your views?
5Evan Hubinger2mo
I would be much happier with that; I think that's much more correct. Then, my objection would just be that at least the sort of counting arguments for deceptive alignment that I like are and always have been about parameterizations rather than functions. I agree that if you try to run a counting argument directly in function space it won't work.
3Ryan Greenblatt2mo
See also discussion here.
-1Alex Turner2mo
How can this be true, when you e.g. say there's "only one saint"? That doesn't make any sense with parameterizations due to internal invariances; there are uncountably many "saints" in parameter-space (insofar as I accept that frame, which I don't really but that's not the point here). I'd expect you to raise that as an obvious point in worlds where this really was about parameterizations. And, as you've elsewhere noted, we don't know enough about parameterizations to make counting arguments over them. So how are you doing that?
3Evan Hubinger2mo
Because it was the transcript of a talk? I was trying to explain an argument at a very high level. And there's certainly not uncountably many; in the infinite bitstring case there would be countably many, though usually I prefer priors that put caps on total computation such that there are only finitely many. I don't really appreciate the psychoanalysis here. I told you what I thought and think, and I have far more evidence about that than you do. As I've said, I usually try to take whatever the most realistic prior is that we can reason about at a high-level, e.g. a circuit prior or a speed prior.
1Chris_Leong2mo
Nabgure senzr gung zvtug or hfrshy: Gurer'f n qvssrerapr orgjrra gur ahzore bs zngurzngvpny shapgvbaf gung vzcyrzrag n frg bs erdhverzragf naq gur ahzore bs cebtenzf gung vzcyrzrag gur frg bs erdhverzragf. Fvzcyvpvgl vf nobhg gur ynggre, abg gur sbezre. Gur rkvfgrapr bs n ynetr ahzore bs cebtenzf gung cebqhpr gur rknpg fnzr zngurzngvpny shapgvba pbagevohgrf gbjneqf fvzcyvpvgl.

From my perspective reading this post, it read to me like "I didn't understand the counting argument, therefore it doesn't make sense" which is (obviously) not very compelling to me.

I definitely appreciate how it can feel frustrating or bad when you feel that someone isn't properly engaging with your ideas. However, I also feel frustrated by this statement. Your comment seems to have a tone of indignation that Quintin and Nora weren't paying attention to what you wrote. 

I myself expected you to respond to this post with some ML-specific reasoning about simplicity and measure of parameterizations, instead of your speculation about a relationship between the universal measure and inductive biases. I spoke with dozens of people about the ideas in OP's post, and none of them mentioned arguments like the one you gave. I myself have spent years in the space and am also not familiar with this particular argument about bitstrings. 

(EDIT: Having read Ryan's comment, it now seems to me that you have exclusively made a simplicity argument without any counting involved, and an empirical claim about the relationship between description length of a mesa objective and the probability of... (read more)

I myself expected you to respond to this post with some ML-specific reasoning about simplicity and measure of parameterizations, instead of your speculation about a relationship between the universal measure and inductive biases. I spoke with dozens of people about the ideas in OP's post, and none of them mentioned arguments like the one you gave. I myself have spent years in the space and am also not familiar with this particular argument about bitstrings.

That probably would have been my objection had the reasoning about priors in this post been sound, but since the reasoning was unsound, I turned to the formalism to try to show why it's unsound.

If these are your real reasons for expecting deceptive alignment, that's fine, but I think you've mentioned this rather infrequently.

I think you're misunderstanding the nature of my objection. It's not that Solomonoff induction is my real reason for believing in deceptive alignment or something, it's that the reasoning in this post is mathematically unsound, and I'm using the formalism to show why. If I weren't responding to this post specifically, I probably wouldn't have brought up Solomonoff induction at all.

This yields a perfe

... (read more)
4Alex Turner2mo
1. This is basically my position as well 2. The cited argument is a counting argument over the space of functions which achieve zero/low training loss.  Indeed, this is a crucial point that I think the post is trying to make. The cited counting arguments are counting functions instead of parameterizations. That's the mistake (or, at least "a" mistake). I'm glad we agree it's a mistake, but then I'm confused why you think that part of the post is unsound.  (Rereads) Rereading the portion in question now, it seems that they changed it a lot since the draft. Personally, I think their argumentation is now weaker than it was before. The original argumentation clearly explained the mistake of counting functions instead of parameterizations, while the present post does not. It instead abstracts it as "an indifference principle", where the reader has to do the work to realize that indifference over functions is inappropriate. 
6Ryan Greenblatt2mo
In this argument, you've implicitly assumed that there is only one function/structure which suffices for being getting high enough training performance to be selected while also not being a long term objective (aka a deceptive objective). I could imagine this being basically right, but it certainly seems non-obvious to me. E.g., there might be many things which are extremely highly correlated with reward that are represented in the world model. Or more generally, there are in principle many objective computations that result in trying as hard to get reward as the deceptive model would try. (The potential for "multiple" objectives only makes a constant factor difference, but this is exactly the same as the case for deceptive objectives.) The fact that these objectives generalize differently maybe implies they aren't "aligned", but in that case there is another key category of objectives: non-exactly-aligned and non-deceptive objectives. And obviously our AI isn't going to be literally exactly aligned. Note that non-exactly-aligned and non-deceptive objectives could suffice for safety in practice even if not perfectly aligned (e.g. due to myopia).
1Evan Hubinger2mo
Yep, that's exactly right. As always, once you start making more complex assumptions, things get more and more complicated, and it starts to get harder to model things in nice concrete mathematical terms. I would defend the value of having actual concrete mathematical models here—I think it's super easy to confuse yourself in this domain if you aren't doing that (e.g. as I think the confused reasoning about counting arguments in this post demonstrates). So I like having really concrete models, but only in the "all models are wrong, but some are useful" sense, as I talk about in "In defense of probably wrong mechanistic models." Also, the main point I was trying to make is that the counting argument is both sound and consistent with known generalization properties of machine learning (and in fact predicts them), and for that purpose I went with the simplest possible formalization of the counting argument.
6Ryan Greenblatt2mo
I found the explanation at the point where you introduce b confusing. Here's a revised version of the text there that would have been less confusing to me (assuming I haven't made any errors):
4Evan Hubinger2mo
Yep, I endorse that text as being equivalent to what I wrote; sorry if my language was a bit confusing.
3Signer2mo
Under this picture, or any other simplicity bias, why NNs with more parameters generalize better?

Paradoxically, I think larger neural networks are more simplicity-biased.

The idea is that when you make your network larger, you increase the size of the search space and thus the number of algorithms that you're considering to include algorithms which take more computation. That reduces the relative importance of the speed prior, but increases the relative importance of the simplicity prior, because your inductive biases are still selecting from among those algorithms according to the simplest pattern that fits the data, such that you get good generalization—and in fact even better generalization because now the space of algorithms in which you're searching for the simplest one in is even larger.

Another way to think about this: if you really believe Occam's razor, then any learning algorithm generalizes exactly to the extent that it approximates a simplicity prior—thus, since we know neural networks generalize better as they get larger, they must be approximating a simplicity prior better as they do so.

This isn't a proper response to the post, but since I've occasionally used counting-style arguments in the past I think I should at least lay out some basic agree/disagree points. So:

  • This post basically-correctly refutes a kinda-mediocre (though relatively-commonly-presented) version of the counting argument.
  • There does exist a version of the counting argument which basically works.
  • The version which works routes through compression and/or singular learning theory.
  • In particular, that version would talk about "goal-slots" (i.e. general-purpose search) showing up for exactly the same reasons that neural networks are able to generalize in the overparameterized regime more generally. In other words, if you take the "counting argument for overfitting" from the post, walk through the standard singular-learning-theory-style response to that story, and then translate that response over to general-purpose search as a specific instance of compression, then you basically get the good version of the counting argument.
... (read more)

Since there are “more” possible schemers than non-schemers, the argument goes, we should expect training to produce schemers most of the time. In Carlsmith’s words:

It's important to note that the exact counting argument you quote isn't one that Carlsmith endorses, just one that he is explaning. And in fact Carlsmith specifically notes that you can't just apply something like the principle of indifference without more reasoning about the actual neural network prior.

(You mention this later in the "simplicity arguments" section, but I think this objection is sufficiently important and sufficiently missing early in the post that it is important to emphasize.)

Quoting somewhat more context:

I start, in section 4.2, with what I call the “counting argument.” It runs as follows:

  1. The non-schemer model classes, here, require fairly specific goals in order to get high reward.
  2. By contrast, the schemer model class is compatible with a very wide range of (beyond- episode) goals, while still getting high reward (at least if we assume that the other require- ments for scheming to make sense as an instrumental strategy are in place—e.g., that the classic goal-guarding story, or some alternative
... (read more)

We argue against the counting argument in general (more specifically, against the presumption of a uniform prior as a "safe default" to adopt in the absence of better information). This applies to the hazy counting argument as well. 

We also don't really think there's that much difference between the structure of the hazy argument and the strict one. Both are trying to introduce some form of ~uniformish prior over the outputs of a stochastic AI generating process. The strict counting argument at least has the virtue of being precise about which stochastic processes it's talking about. 

If anything, having more moving parts in the causal graph responsible for producing the distribution over AI goals should make you more skeptical of assigning a uniform prior to that distribution. 

I agree that you can't adopt a uniform prior. (By uniform prior, I assume you mean something like, we represent goals as functions from world states to a (real) number where the number says how good the world state is, then we take a uniform distribution over this function space. (Uniform sampling from function space is extremely, extremely cursed for analysis related reasons without imposing some additional constraints, so it's not clear uniform sampling even makes sense!))

Separately, I'm also skeptical that any serious historical arguments were actually assuming a uniform prior as opposed to trying to actual reason about the complexity/measure of various goal in terms of some fixed world model given some vague guess about the representation of this world model. This is also somewhat dubious due to assuming a goal slot, assuming a world model, and needing to guess at the representation of the world model.

(You'll note that ~all prior arguements mention terms like "complexity" and "bits".)

Of course, the "Against goal realism" and "Simplicity arguments" sections can apply here and indeed, I'm much more sympathetic to these sections than to the counting argument section which seems like a strawman as far as I can tell. (I tried to get to ground on this by communicating back and forth some with you and some with Alex Turner, but I failed, so now I'm just voicing my issues for third parties.)

I don't think this is a strawman. E.g., in How likely is deceptive alignment?, Evan Hubinger says:

We're going to start with simplicity. Simplicity is about specifying the thing that you want in the space of all possible things. You can think about simplicity as “How much do you have to aim to hit the exact thing in the space of all possible models?” How many bits does it take to find the thing that you want in the model space? And so, as a first pass, we can understand simplicity by doing a counting argument, which is just asking, how many models are in each model class?

First, how many Christs are there? Well, I think there's essentially only one, since there's only one way for humans to be structured in exactly the same way as God. God has a particular internal structure that determines exactly the things that God wants and the way that God works, and there's really only one way to port that structure over and make the unique human that wants exactly the same stuff.

Okay, how many Martin Luthers are there? Well, there's actually more than one Martin Luther (contrary to actual history) because the Martin Luthers can point to the Bible in different ways. There's a lot of different eq

... (read more)
5Ryan Greenblatt2mo
I'm sympathetic to pushing back on counting arguments on the ground 'it's hard to know what the exact measure should be, so maybe the measure on the goal of "directly pursue high performance/anything nearly perfectly correlated the outcome that it reinforced (aka reward)" is comparable/bigger than the measure on "literally any long run outcome"'. So I appreciate the push back here. I just think the exact argument and the comparison to overfitting is a strawman. (Note that above I'm assuming a specific goal slot, that the AI's predictions are aware of what its goal slot contains, and that in order for the AI to perform sufficiently well as to be a plausible result of training it has to explicitly "play the training game" (e.g. explicitly reason about and try to get high performance). It also seems reasonable to contest these assumption, but this is a different thing than the counting argument.) (Also, if we imagine an RL'd neural network computing a bunch of predictions, then it does seem plausible that it will have a bunch of long horizon predictions with higher aggregate measure than predicting things that perfectly correlate with the outcome that was reinforced (aka reward)! As in, if we imagine randomly sampling a linear probe, it will be far more likely to sample a probe where most of the variance is driven by long run outcomes than to sample a linear probe which is almost perfectly correlated with reward (e.g. a near perfect predictor of reward up to monotone regression). Neural networks are likely to compute a bunch of long range predictions at least as intermediates, but they only need to compute things that nearly perfectly correlate with reward once! (With some important caveats about transfer from other distributions.)) I also think Evan's arguments are pretty sloppy in this presentation and he makes a bunch of object level errors/egregious simplifications FWIW, but he is actually trying to talk about models represented in weight space and how many bit
4Ryan Greenblatt2mo
[Low importance aside] I think this is equivalent to a well known approximation from algorithmic information theory. I think this approximation might be too lossy in practice in the case of actual neural nets though.

Deep learning is strongly biased toward networks that generalize the way humans want— otherwise, it wouldn’t be economically useful.

This is NOT what the evidence supports, and super misleadingly phrased. (Either that, or it's straightup magical thinking, which is worse)

The inductive biases / simplicity biases of deep learning are poorly understood but they almost certainly don't have anything to do with what humans want, per se. (that would be basically magic) Rather, humans have gotten decent at intuiting them, such that humans can often predict how the neural network will generalize in response to such-and-such training data. i.e. human intuitive sense of simplicity is different, but not totally different, at least not always, from the actual simplicity biases at play.

Stylized abstract example: Our current AI is not generalizing in the way we wanted it to. Looking at its behavior, and our dataset, we intuit that the dataset D is narrow/nondiverse in ways Y and Z and that this could be causing the problem; we go collect more data so that our dataset is diverse in those ways, and try again, and this time it works (i.e. the AI generalizes to unseen data X). Why did this happen? Why ... (read more)

2Alex Turner2mo
Seems like a misunderstanding. It seems to me that you are alleging that Nora/Quintin believe there is a causal arrow from "Humans want X generalization" to "NNs have X generalization"? If so, I think that's an uncharitable reading of the quoted text.
4Daniel Kokotajlo2mo
I said "Either that, or it's straightup magical thinking" which was referring to the causal arrow hypothesis. I agree it's unlikely that they would endorse the causal arrow / magical thinking hypothesis, especially once it's spelled out like that.  What do you think they meant by "Deep learning is strongly biased toward networks that generalize the way humans want— otherwise, it wouldn’t be economically useful?"
4Alex Turner2mo
I think they meant that there is an evidential update from "it's economically useful" upwards on "this way of doing things tends to produce human-desired generalization in general and not just in the specific tasks examined so far."  Perhaps it's easy to consider the same style of reasoning via: "The routes I take home from work are strongly biased towards being short, otherwise I wouldn't have taken them home from work."
2Daniel Kokotajlo2mo
Thanks. The routes-home example checks out IMO. Here's another one that also seems to check out, which perhaps illustrates why I feel like the original claim is misleading/unhelpful/etc.: "The laws of ballistics strongly bias aerial projectiles towards landing on targets humans wanted to hit; otherwise, ranged weaponry wouldn't be militarily useful." There's a non-misleading version of this which I'd recommend saying instead, which is something like "Look we understand the laws of physics well enough and have played around with projectiles enough in practice, that we can reasonably well predict where they'll land in a variety of situations, and design+aim weapons accordingly; if this wasn't true then ranged weaponry wouldn't be militarily useful." And I would endorse the corresponding claim for deep learning: "We understand how deep learning networks generalize well enough, and have played around with them enough in practice, that we can reasonably well predict how they'll behave in a variety of situations, and design training environments accordingly; if this wasn't true then deep learning wouldn't be economically useful." (To which I'd reply "Yep and my current understanding of how they'll behave in certain future scenarios is that they'll powerseek, for reasons which others have explained... I have some ideas for other, different training environments that probably wouldn't result in undesired behavior, but all of this is still pretty up in the air tbh I don't think anyone really understands what they are doing here nearly as well as e.g. cannoneers in 1850 understood what they were doing.")
2Daniel Kokotajlo2mo
To put it in terms of the analogy you chose: I agree (in a sense) that the routes you take home from work are strongly biased towards being short, otherwise you wouldn't have taken them home from work. But if you tell me that today you are going to try out a new route, and you describe it to me and it seems to me that it's probably going to be super long, and I object and say it seems like it'll be super long for reasons XYZ, it's not a valid reply for you to say "don't worry, the routes I take home from work are strongly biased towards being short, otherwise I wouldn't take them." At least, it seems like a pretty confusing and maybe misleading thing to say. I would accept "Trust me on this, I know what I'm doing, I've got lots of experience finding short routes" I guess, though only half credit for that since it still wouldn't be an object level reply to the reasons XYZ and in the absence of such a substantive reply I'd start to doubt your expertise and/or doubt that you were applying it correctly here (especially if I had an error theory for why you might be motivated to think that this route would be short even if it wasn't.)

I think that if you do assume a fixed goal slot and outline an overall architecture, then there are pretty good arguments for a serious probabilty of scheming.

(Though there are also plenty of bad arguments, including some that people have made in the past : ).)

That said, I'm sympathetic to some version of the "Against goal realism" argument applying to models which are sufficiently useful. As in, the first transformatively useful models won't in practice contain have internal (opaque-to-human-overseers) goals such that the traditional story for scheming doesn't apply.

(However, it's worth noting that at least some humans do seem to have internal goals and reductionism doesn't defeat this intuition. It's not super clear that the situation with humans is well described as a "goal slot", though there is pretty clearly some stuff that could be changed in a human brain that would cause them to be well described as coherantly pursue different goals. So arguing that AIs won't have internal goals in a way that could result in scheming does require noting some ways in which you're argument doesn't apply to humans. More strongly, humans can and do scheme even in cases where some overseer sele... (read more)

(I might write a longer response later, but I thought it would be worth writing a quick response now. Cross-posted from the EA forum, and I know you've replied there, but I'm posting anyway.)

I have a few points of agreement and a few points of disagreement:

Agreements:

  • The strict counting argument seems very weak as an argument for scheming, essentially for the reason you identified: it relies on a uniform prior over AI goals, which seems like a really bad model of the situation.
  • The hazy counting argument—while stronger than the strict counting argument—still seems like weak evidence for scheming. One way of seeing this is, as you pointed out, to show that essentially identical arguments can be applied to deep learning in different contexts that nonetheless contradict empirical evidence.

Some points of disagreement:

  • I think the title overstates the strength of the conclusion. The hazy counting argument seems weak to me but I don't think it's literally "no evidence" for the claim here: that future AIs will scheme.
  • I disagree with the bottom-line conclusion: "we should assign very low credence to the spontaneous emergence of scheming in future AI systems—perhaps 0.1% or less"
    • I think it's
... (read more)
6Ryan Greenblatt2mo
It's worth noting here that Carlsmith's original usage of the term scheming just refers to AIs that perform well on training and evaluations for instrumental reasons because they have longer run goals or similar. So, AIs lying because this was directly reinforced wouldn't itself be scheming behavior in Carlsmith's terminology. However, it's worth noting that part of Carlsmith's argument involves arguing that smart AIs will likely have to explicitly reason about the reinforcement process (sometimes called playing the training game) and this will likely involve lying.
5Matthew Barnett2mo
Perhaps I was being too loose with my language, and it's possible this is a pointless pedantic discussion about terminology, but I think I was still pointing to what Carlsmith called schemers in that quote. Here's Joe Carlsmith's terminological breakdown: The key distinction in my view is whether the designers of the reward function intended for lies to be reinforced or not. [ETA: this was confusingly stated. What I meant is that if a people design a reward function that accidentally reinforces lying in order to obtain power, it seems reasonable to call the agent that results from training on that reward function a "schemer" given Carlsmith's terminology, and common sense.] If lying to obtain power is reinforced but the designers either do not know this, or do not know how to mitigate this behavior, then it still seems reasonable to call the resulting model a "schemer". In Ajeya Cotra's story, for example: 1. Alex was incentivized to lie because it got rewards for taking actions that were superficially rated as good even if they weren't actually good, i.e. Alex was "lying because this was directly reinforced". She wrote, "Because humans have systematic errors in judgment, there are many scenarios where acting deceitfully causes humans to reward Alex’s behavior more highly. Because Alex is a skilled, situationally aware, creative planner, it will understand this; because Alex’s training pushes it to maximize its expected reward, it will be pushed to act on this understanding and behave deceptively." 2. Alex was "playing the training game", as Ajeya Cotra says this explicitly several times in her story. 3. Alex was playing the training game in order to get power for itself or for other AIs; clearly, as the model literally takes over the world and disempowers humanity at the end. 4. Alex kind of didn't appear to purely care about reward-on-the-episode, since it took over the world? Yes, Alex cared about rewards, but not necessarily on this episode. Maybe I'm wro
4Ryan Greenblatt2mo
Hmm, I don't think the intention is the key thing (at least with how I use the word and how I think Joe uses the word), I think the key thing is whether the reinforcement/reward process actively incentivizes bad behavior. Overall, I use the term to mean basically the same thing as "deceptive alignment". (But more specifically pointing the definition in Joe's report which depends less on some notion of mesa-optimization and is a bit more precise IMO.)
1Matthew Barnett2mo
I confusingly stated my point (and retracted my specific claim in the comment above). I think the rest of my comment basically holds, though. Here's what I think is a clearer argument: * The term "schemer" evokes an image of someone who is lying to obtain power. It doesn't particularly evoke a backstory for why the person became a liar in the first place. * There are at least two ways that AIs could arise that lie in order to obtain power: * The reward function could directly reinforce the behavior of lying to obtain power, at least at some point in the training process. * The reward function could have no defects (in the sense of not directly reinforcing harmful behavior), and yet an agent could nonetheless arise during training that lies in order to obtain power, simply because it is a misaligned inner optimizer (broadly speaking) * In both cases, one can imagine the AI eventually "playing the training game", in the sense of having a complete understanding of its training process and deliberately choosing actions that yield high reward, according to its understanding of the training process * Since both types of AIs are: (1) playing the training game, (2) lying in order to obtain power, it makes sense to call both of them "schemers", as that simply matches the way the term is typically used.  For example, Nora and Quintin started their post with, "AI doom scenarios often suppose that future AIs will engage in scheming— planning to escape, gain power, and pursue ulterior motives, while deceiving us into thinking they are aligned with our interests." This usage did not specify the reason for the deceptive behavior arising in the first place, only that the behavior was both deceptive and aimed at gaining power. * Separately, I am currently confused at what it means for a behavior to be "directly reinforced" by a reward function, so I'm not completely confident in these arguments, or my own line of reasoning here. My best guess is that these are fuz
2Ryan Greenblatt2mo
I agree this matches typical usage (and also matches usage in the overall post we're commenting on), but sadly the word schemer in the context of Joe's report means something more specific. I'm sad about the overall terminology situation here. It's possible I should just always use a term like beyond-episode-goal-style-scheming. I agree this distinction is fuzzy, but I think there is likely to be an important distinction because the case where the behavior isn't due to things well described as beyond-episode-goals, it should be much easier to study. See here for more commentary. There will of course be a spectrum here.
4Ryan Greenblatt2mo
I think in Ajeya's story the core threat model isn't well described as scheming and is better described as seeking some proxy of reward.
1Alex Turner2mo
I agree, they're wrong to claim it's "no evidence." I think that counting arguments are extremely slight evidence against scheming, because they're weaker than the arguments I'd expect our community's thinkers to find in worlds where scheming was real. (Although I agree that on the object-level and in isolation, the arguments are tiiiny positive evidence.)

The current literature on scheming appears to have been inspired by Paul Christiano’s speculations about malign intelligences in Solomonoff induction

This doesn't seem right. The linked post by Paul here is about the (extremely speculative) case where consequentialist life emerges organically inside of full blown simulations (e.g. evolving from scratch) while arguments about ML models never go here.

Regardless, concerns and arguments about scheming are much older than Paul's posts on this topic.

(That said, I do think that people have made scheming style arguments based on intuitions from thinking about AIXI and the space of turing machines at various points. Though this was never very key and I don't believe these arguments are ever in reference to cases where a literal simulation evolves life.)

In reality, the problem faced by evolution and by SGD is much easier than this: producing systems that behave the right way in all scenarios they are likely to encounter. In virtue of their aligned behavior, these systems will be “aimed at the right things” in every sense that matters in practice.

I find this passage remarkable, given that so many people are choosing to to have few or no children that fertility has fallen to 0.78 in Korea and 1.0 in China. Presumably you're aware of these (or similar) facts and intended the meaning of this passage to be compatible with them, but I'm having trouble figuring out how...

By contrast, goal realism leads only to unfalsifiable speculation about an “inner actress” with utterly alien motivations.

In order for such speculation to be unfalsifiable, it seemingly has to be the case that we're unable to ever develop good enough interpretability tools to definitively say whether the AI in question has such internal motivations. This could well turn out to be true, but I don't understand how you're able to predict this now. (Or maybe you mean something else by "unfalsifiable" but I can't see what it could be. ETA: Maybe you mean "unfalsifiable... (read more)

3Nora Belrose2mo
The point of that section is that "goals" are not ontologically fundamental entities with precise contents, in fact they could not possibly be so given a naturalistic worldview. So you don't need to "target the inner search," you just need to get the system to act the way you want in all the relevant scenarios. The modern world is not a relevant scenario for evolution. "Evolution" did not need to, was not "intending to," and could not have designed human brains so that they would do high inclusive genetic fitness stuff even when the environment wildly dramatically changes and culture becomes completely different from the ancestral environment.

I feel like there's a somewhat common argument about RL not being all that dangerous because it generalizes the training distribution cautiously - being outside the training distribution isn't going to suddenly cause an RL system to make multi-step plans that are implied but never seen in the training distribution, it'll probably just fall back on familiar, safe behavior.

To me, these arguments feel like they treat present-day model-free RL as the "central case," and model-based RL as a small correction.

Anyhow, good post, I like most of the arguments, I just felt my reaction to this particular one could be made in meme format.

  • It seems like there are “lots of ways” that a model could end up massively overfitting and still get high training performance.
  • So absent some additional story about why training won’t select an overfitter, it feels like the possibility should be getting substantive weight.

FWIW, once I learned more about the problem of induction, I realized that there do exist additional stories explaining why training won't select an overfitter. Or perhaps to put it differently, after I understood the problem of induction better it no longer seemed to me that there were lots of ways a model could massively overfit and still get high training performance. (That is, it seems to me there are many MORE ways it could not overfit)

1Nora Belrose2mo
It depends what you mean by a "way" the model can overfit. Really we need to bring in measure theory to rigorously talk about this, and an early draft of this post actually did introduce some measure-theoretic concepts. Basically we need to define: * What set are we talking about, * What measure we're using over that set, * And how that measure relates to the probability measure over possible AIs. The English locution "lots of ways to do X" can be formalized as "the measure of X-networks is high." And that's going to be an empirical claim that we can actually debate.
2Daniel Kokotajlo2mo
I think I mean the same thing you do? "The measure of X-networks is high."
1Nora Belrose2mo
With respect to which measure though? You have to define a measure, there are going to be infinitely many possible measures you could define on this space. And then we'll have to debate if your measure is a good one.
4Daniel Kokotajlo2mo
The actual measure that nature uses to determine the model weights at the end of training -- taking into account the random initialization and also the training process. I'm talking about the (not-yet-fully-understood) inductive biases of neural networks in practice.
2Daniel Kokotajlo2mo
Added clarification: When I said "once I understood the problem of induction better" I was referring specifically to the insight evhub attempts to convey with his example about infinite bitstrings. Simpler circuits, policies, goals, strategies, whatever can be instantiated in more ways than all their complex alternatives combined.
4Nora Belrose2mo
I think the infinite bitstring case has zero relevance to deep learning. There does exist a concept you might call "simplicity" which is relevant to deep learning. The neural network Gaussian process describes the prior distribution over functions which is induced by the initialization distribution over neural net parameters. Under weak assumptions about the activation function and initialization variance, the NNGP is biased toward lower frequency functions. I think this cuts against scheming, and we plan to write up a post on this in the next month or two.
2Evan Hubinger2mo
I think you are still not really understanding my objection. It's not that there is a "finite bitstring case" and an "infinite bitstring case". My objection is that the sort of finite bitstring analysis that you use does not yield any well-defined mathematical object that you could call a prior, and certainly not one that would predict generalization.
1Nora Belrose2mo
I never used any kind of bitstring analysis.
2Evan Hubinger2mo
Yes, that's exactly the problem: you tried to make a counting argument, but because you didn't engage with the proper formalism, you ended up using reasoning that doesn't actually correspond to any well-defined mathematical object. Analogously, it's like you wrote an essay about why 0.999... != 1 and your response to "under the formalism of real numbers as Dedekind cuts, those are identical" was "where did I say I was referring to Dedekind cuts?" It's fine if you don't want to use the standard formalism, but you need some formalism to anchor your words to, otherwise you're just pushing around words with no real way to ensure that your words actually correspond to something. I think the 0.999... != 1 analogy is quite apt here, because the problem really is that there is no formalism under which 0.999... != 1 that looks anything like the real numbers that you know, in the same way that there really is no formalism under which the sort of reasoning that you're using is meaningful.
4Alex Turner2mo
No. I think you are wrong. This passage makes me suspect that you didn't understand the arguments Nora was trying to make. Her arguments are easily formalizable as critiquing an indifference principle over functions in function-space, as opposed to over parameterizations in parameter-space. I'll write this out for you if you really want me to. I think you should be more cautious at unilaterally diagnosing Nora's "errors", as opposed to asking for clarification, because I think you two agree a lot more than you realize.
2Evan Hubinger2mo
I agree that there is a valid argument that critiques counting arguments over function space that sort of has the same shape as the one presented in this post. If that was what the authors had in mind, it was not what I got from reading the post, and I haven't seen anyone making that clarification other than yourself. Regardless, though, I think that's still not a great objection to counting arguments for deceptive alignment in general, because it's explicitly responding only to a very weak and obviously wrong form of a counting argument. My response there is just that of course you shouldn't run a counting argument over function space—I would never suggest that.
1Alex Turner2mo
I think you should have asked for clarification before making blistering critiques about how Nora "ended up using reasoning that doesn't actually correspond to any well-defined mathematical object." I think your comments paint a highly uncharitable and (more importantly) incorrect view of N/Q's claims. Your presentations often include a counting argument over a function space, in the form of "saints" versus "schemers" and "sycophants." So it seems to me that you do suggest that. What am I missing? I also welcome links to counting arguments which you consider stronger. I know you said you haven't written one up yet to your satisfaction, but surely there have to be some non-obviously wrong and weak arguments written up, right?
4Evan Hubinger2mo
I'm happy to apologize if I misinterpreted anyone, but afaict my critique remains valid. My criticism is precisely that counting arguments over function space aren't generally well-defined, and even if they were they wouldn't be the right way to run a counting argument. So my criticism that the original post misunderstands how to properly run a counting argument still seems correct to me. Perhaps you could say that it's not the authors' fault, that they were responding to weak arguments that other people were actually making, but regardless the point remains that the authors haven't engaged with the sort of counting arguments that I actually think are valid. What makes you think that's intended to be a counting argument over function space? I usually think of this as a counting argument over infinite bitstrings, as I noted in my comment (though there are many other valid presentations). It's possible I said something in that talk that gave a misleading impression there, but I certainly don't believe and have never believed in any counting arguments over function space.
2Alex Turner2mo
Going back through the post, Nora+Quintin indeed made a specific and perfectly formalizable claim here:  They're making a perfectly valid point. The point was in the original post AFAICT -- it wasn't just only now explained by me. I agree that they could have presented it more clearly, but that's a way different critique than you're "using reasoning that doesn't actually correspond to any well-defined mathematical object." If that's truly your remaining objection, then I think that you should retract the unmerited criticisms about how they're trying to prove 0.9999... != 1 or whatever. In my opinion, you have confidently misrepresented their arguments, and the discussion would benefit from your revisions.   And then it'd be nice if someone would provide links to the supposed valid counting arguments! From my perspective, it's very frustrating to hear that there (apparently) are valid counting arguments but also they aren't the obvious well-known ones that everyone seems to talk about. (But also the real arguments aren't linkable.) If that's truly the state of the evidence, then I'm happy to just conclude that Nora+Quintin are right, and update if/when actually valid arguments come along.

If that's truly your remaining objection, then I think that you should retract the unmerited criticisms about how they're trying to prove 0.9999... != 1 or whatever. In my opinion, you have confidently misrepresented their arguments, and the discussion would benefit from your revisions.

This point seems right to me: if the post is specifically about representable functions than that is a valid formalization AFAICT. (Though a extremely cursed formalization for reasons mentioned in a variety of places. And if you dropped "representable", then it's extremely, extremely cursed for various analysis related reasons, though I think there is still a theoretically sound uniform measure maybe???)

It would also be nice if the original post:

  • Clarified that the rebuttal is specifically about a version of the counting-argument which counts functions.
  • Noted that people making counting arguments weren't intending to count functions, though this might be a common misconception about counting arguments. (Seems fine to also clarify that existing counting arguments are too hand wavy to really engage with if that's the view also.) (See also here.)
5Ryan Greenblatt2mo
Personally, I don't think there are "solid" counting arguments, but I think you can think though a bunch more cases and feel like the underlying intuition is at least somewhat reasonable. Overall, I'm a simple man, I still like Joe's report : ). Fair enough if you don't find the arguments in here convincing. I think Joe's report is pretty close to the SOTA with open mindedness and a bit of reinvention work to fill in various gaps.
1Nora Belrose2mo
I definitely thought you were making a counting argument over function space, and AFAICT Joe also thought this in his report. The bitstring version of the argument, to the extent I can understand it, just seems even worse to me. You're making an argument about one type of learning procedure, Solomonoff induction, which is physically unrealizable and AFAICT has not even inspired any serious real-world approximations, and then assuming that somehow the conclusions will transfer over to a mechanistically very different learning procedure, gradient descent. The same goes for the circuit prior thing (although FWIW I think you're very likely wrong that minimal circuits can be deceptive).

I definitely thought you were making a counting argument over function space

I've argued multiple times that Evan was not intending to make a counting argument in function space:

  • In discussion with Alex Turner (TurnTrout) when commenting on an earlier draft of this post.
  • In discussion with Quintin after sharing some comments on the draft. (Also shared with you TBC.)
  • In this earlier comment.

(Fair enough if you never read any of these comments.)

As I've noted in all of these comments, people consistently use terminology when making counting style arguments (except perhaps in Joe's report) which rules out the person intending the argument to be about function space. (E.g., people say things like "bits" and "complexity in terms of the world model".)

(I also think these written up arguments (Evan's talk in particular) are very hand wavy, and just provide a vague intuition. So regardless of what he was intending, the actual words of the argument aren't very solid IMO. Further, using words that rule out the intention of function space doesn't necessarily imply there is an actually good model behind these words. To actually get anywhere with this reasoning, I think you'd have to reinven... (read more)

3Alex Turner2mo
Aren't these arguments about simplicity, not counting? 
2Evan Hubinger2mo
Sorry about that—I wish you had been at the talk and could have asked a question about this. I agree that Solomonoff induction is obviously wrong in many ways, which is why you want to substitute it out for whatever the prior is that you think is closest to deep learning that you can still reason about theoretically. But that should never lead you to do a counting argument over function space, since that is never a sound thing to do.
1Alex Turner2mo
Do you agree that "instrumental convergence -> meaningful evidence for doom" is also unsound, because it's a counting argument that most functions of shape Y have undesirable property X?
3Evan Hubinger2mo
I think instrumental convergence does provide meaningful evidence of doom, and you can make a valid counting argument for it, but as with deceptive alignment you have to run the counting argument over algorithms not over functions.
-8Nora Belrose2mo
1Nora Belrose2mo
I obviously don't think the counting argument for overfitting is actually sound, that's the whole point. But I think the counting argument for scheming is just as obviously invalid, and misuses formalisms just as egregiously, if not moreso. I deny that your Kolmogorov framework is anything like "the proper formalism" for neural networks. I also deny that the counting argument for overfitting is appropriately characterized as a "finite bitstring" argument, because that suggests I'm talking about Turing machine programs of finite length, which I'm not- I'm directly enumerating functions over a subset of the natural numbers. Are you saying the set of functions over 1...10,000 is not a well defined mathematical object?

I obviously don't think the counting argument for overfitting is actually sound, that's the whole point.

Yes, I'm well aware. The problem is that when you make the counting argument for overfitting, you do so in a way that seriously misuses the formalism, which is why the argument fails. So you can't draw any lessons about counting arguments for deception from the failure of your counting argument for overfitting.

But I think the counting argument for scheming is just as obviously invalid, and misuses formalisms just as egregiously, if not moreso.

Then show me how! If you think there are errors in the math, please point them out.

Of course, it's worth stating that I certainly don't have some sort of airtight mathematical argument proving that deception is likely in neural networks—there are lots of assumptions there that could very well be wrong. But I do think that the basic style of reasoning employed by such arguments is sound.

I deny that your Kolmogorov framework is anything like "the proper formalism" for neural networks.

Err... I'm using K-complexity here because it's a simple framework to reason about, but my criticism isn't "you should use K-complexity to reason about... (read more)

0Nora Belrose2mo
I'm not aware of any actual math behind the counting argument for scheming. I've only ever seen handwavy informal arguments about the number of Christs vs Martin Luthers vs Blaise Pascals. There certainly was no formal argument presented in Joe's extensive scheming report, which I assumed would be sufficient context for writing this essay.
5Evan Hubinger2mo
Well, I presented a very simple formulation in my comment, so that could be a reasonable starting point. But I agree that unfortunately there hasn't been that much good formal analysis here that's been written up. At least on my end, that's for two reasons: 1. Most of the formal analysis of this form that I've published (e.g. this and this) has been focused on sycophancy (human imitator vs. direct translator) rather than deceptive alignment, as sycophancy is a substantially more tractable problem. Finding a prior that reasonably rules out deceptive alignment seems quite out of reach to me currently; at one point I thought a circuit prior might do it, but I now think that circuit priors don't get rid of deceptive alignment. 2. I'm currently more optimistic about empirical evidence rather than theoretical evidence for resolving this question, which is why I've been focusing on projects such as Sleeper Agents.
-6Nora Belrose2mo

the problem faced by evolution and by SGD is much easier than this: producing systems that behave the right way in all scenarios they are likely to encounter.

I think you mean "in all scenarios they are likely to encounter *on the training distribution* / in the ancestral environment right? That's importantly different.

I wrote up my views on the principle of indifference here:

https://www.lesswrong.com/posts/3PXBK2an9dcRoNoid/on-having-no-clue

I agree that it has certain philosophical issues, but I don’t believe that this is as fatal to counting arguments as you believe.

Towards the end I write:

“The problem is that we are making an assumption, but rather than owning it, we're trying to deny that we're making any assumption at all, ie. "I'm not assuming a priori A and B have equal probability based on my subjective judgement, I'm using the principle of indifference". Roll to... (read more)

In favour of goal realism

Suppose your looking at an AI that is currently placed in a game of chess. 

It has a variety of behaviours. It moves pawns forward in some circumstances. It takes a knight with a bishop in a different circumstance. 

You could describe the actions of this AI by producing a giant table of "behaviours". Bishop taking behaviours in this circumstance. Castling behaviour in that circumstance. ... 

But there is a more compact way to represent similar predictions. You can say it's trying to win at chess. 

The "trying to win... (read more)

We can salvage a counting argument. But it needs to be a little subtle. And it's all about the comments, not the code.

Suppose a neural network has 1 megabyte of memory. To slightly oversimplify, lets say it can represent a python file of 1 megabyte. 

One option is for the network to store a giant lookup table. Lets say the network needs half a megabyte to store the training data in this table. This leaves the other half free to be any rubbish. Hence around  possible networks.

The other option is for the network to implement a simple ... (read more)

Despite not answering all possible goal-related questions a priori, the reductionist perspective does provide a tractable research program for improving our understanding of AI goal development. It does this by reducing questions about goals to questions about behaviors observable in the training data.

[emphasis mine]

This might be described as "a reductionist perspective". It is certainly not "the reductionist perspective", since reductionist perspectives need not limit themselves to "behaviors observable in the training data".

A more reasonable-to-my-mind b... (read more)

More generally, John Miller and colleagues have found training performance is an excellent predictor of test performance, even when the test set looks fairly different from the training set, across a wide variety of tasks and architectures.

Seems like figure 1 from Miller et al is a plot of test performance vs. "out of distribution" test performance. One might expect plots of training performance vs. "out of distribution" test performance to have more spread.

1Nora Belrose2mo
I doubt there would be much difference, and I think the alignment-relevant comparison is to compare in-distribution but out-of-sample performance to out-of-distribution performance. We can easily do i.i.d. splits of our data, that's not a problem. You might think it's a problem to directly test the model in scenarios where it could legitimately execute a takeover if it wanted to.
1Donald Hobson1mo
Taking IID samples can be hard actually. Suppose you train an LLM on news articles. And each important real world event has 10 basically identical news articles written about it. Then a random split of the articles will leave the network being tested mostly on the same newsworthy events that were in the training data.  This leaves it passing the test, even if it's hopeless at predicting new events and can only generate new articles about the same events.  When data duplication is extensive, making a meaningful train/test split is hard.  If the data was perfect copy and paste duplicated, that could be filtered out. But often things are rephrased a bit. 

I think this is an excellent post. I really liked the insight about the mechanisms (and mistakes) shared by the counting arguments behind AI doom and behind "deep learning surely won't generalize." Thank you for writing this; these kinds of loose claims have roamed freely for far too long.

EDIT: Actually this post is weaker than a draft I'd read. I still think it's good, but missing some of the key points I liked the most. And I'm not on board with all of the philosophical claims about e.g. generalized objections to the principle of indifference (in part because I don't understand them).