这是用户在 2024-6-15 20:19 为 https://plato.stanford.edu/entries/rule-following/ 保存的双语快照页面,由 沉浸式翻译 提供双语支持。了解如何保存?

Rule-Following and Intentionality
遵循规则和意向

First published Tue Apr 12, 2022
首次发布于 2022 年 4 月 12 日(星期二)

Ludwig Wittgenstein’s reflections on rule-following—principally, sections 138–242 of Philosophical Investigations and section VI of Remarks on the Foundations of Mathematics—raise a series of provoking questions and puzzles about the nature of language and thought. The literature on this topic is vast. We’ll structure our discussion around Saul Kripke’s Wittgenstein on Rules and Private Language (1982), the most widely discussed commentary on Wittgenstein on rule-following.[1] In this book, Kripke’s Wittgenstein famously develops a “sceptical challenge” to the idea that that there are facts about the meanings of linguistic expressions and contents of thoughts, and goes on to propose a “sceptical solution” to the challenge, which attempts to preserve the propriety of talk of meaning and content while conceding to the sceptic the non-existence of the kind of semantic or intentional facts on which she casts doubt. After some preliminary comments about rules, meaning, and content (§1), we’ll outline the sceptical argument (§2), and we’ll offer an overview of some of the main responses to it. First, we’ll discuss sceptical responses of the sort proposed by Kripke’s Wittgenstein himself (§3). Then, we’ll consider straight responses, which advance candidate meaning-constituting facts. We’ll examine in detail the two forms of straight solution that are most widely discussed in the secondary literature, reductive dispositionalism (§4) and non-reductionism (§5).
路德维希·维特根斯坦关于规则遵循的反思主要集中在《哲学调查》的第 138 至 242 段和《数学基础评论》第 VI 节,引发了一系列关于语言和思维本质的挑衅性问题和谜题。关于这一主题的文献非常广泛。我们将围绕索尔·克里普克的《维特根斯坦论规则与私语》(1982 年)展开讨论,这是关于维特根斯坦规则遵循观点最广泛讨论的评论。在这本书中,克里普克的维特根斯坦著名地提出了对语言表达含义和思想内容有关事实的“怀疑挑战”,并进一步提出了对这一挑战的“怀疑解决方案”,试图在承认怀疑者对语义或意图事实的非存在性的同时,保留关于意义和内容的讨论的恰当性。在对规则、意义和内容进行一些初步评论(§1)之后,我们将概述怀疑论证(§2)并提供一些主要对其的回应概览。 首先,我们将讨论克里普克的维特根斯坦本人提出的怀疑性回应(§3)。然后,我们将考虑直接回应,即提出候选的构成意义的事实。我们将详细研究在二级文献中广泛讨论的两种直接解决方案形式,即简化处置主义(§4)和非简化主义(§5)。

For the purposes of this entry, we’ll understand ‘intentionality’ in a broad sense, so that it covers the meanings of linguistic expressions (and utterances) and also the contents of propositional attitudes. Clearly, these notions are closely related: just as we can say of the sentence 68+57=125 that it means that 68 plus 57 is 125, we can say of Octavia’s belief that it has the content that 68 plus 57 is 125 (or that Octavia believes that 68 plus 57 is 125).
为了本条目的目的,我们将广义地理解‘意向性’,使其涵盖语言表达的意义(和话语)以及命题态度的内容。显然,这些概念密切相关:正如我们可以说句子‘ 68+57=125 ’的意思是 68 加 57 等于 125,我们也可以说 Octavia 的信念具有 68 加 57 等于 125 的内容(或 Octavia 相信 68 加 57 等于 125)。

1. Rules, Meaning, and Content
1. 规则、含义和内容

What does the notion of rule-following have to do with the notions of linguistic meaning and mental content? For our purposes, the important point is that meaning something by a linguistic expression is analogous to following a rule. Suppose I write out the beginning of an arithmetical series
规则遵循的概念与语言意义和心理内容的概念有什么关系?对我们来说,重要的一点是,通过语言表达的意思类似于遵循规则。假设我写下一个算术序列的开头

2,4,6,8,10,

If the rule I’m following is add 2, the continuation
如果我遵循的规则是加 2,则继续

12,14,16,

is correct, in that it accords with the rule I’m following, while the continuation
是正确的,因为它符合我遵循的规则,而继续

13,19,20,

is incorrect, in that it fails to accord with the rule I’m following.
错误的,因为它不符合我正在遵循的规则。

We have here an analogy with my meaning something by a linguistic expression. Say that I mean blue by ‘blue’. Then, ‘blue’ is correctly applicable to, for example, a US postbox but not to a ripe (Roma) tomato. We can express this point by saying that the former application accords with the predicate’s meaning while the latter application fails to do so. Given this analogy, arguments about rule-following have consequences for our conception of linguistic meaning: if an argument shows that there are no facts about which rule an agent is following, it may also show that there are no facts about what a speaker means by a linguistic expression.
我们这里有一个类比,类似于我的用语言表达某种意思。假设我用“蓝色”表示蓝色。那么,“蓝色”可以正确适用于例如美国的邮箱,但不适用于成熟的(罗马)番茄。我们可以通过说前者的应用符合谓词的意思,而后者的应用则不符合来表达这一点。根据这个类比,关于遵循规则的论证对我们的语言意义概念有重要影响:如果一个论证显示一个行为者遵循的规则没有任何事实依据,它也可能表明说话者用语言表达的意思没有任何事实依据。

Note that the notion of accord in play in the case of following a rule is also in play in our conception of mental states with intentional content generally. Say that one intends to attend the performance of Sartre’s play No Exit at the Hopewell Theatre on Wednesday. Then, one’s attending the performance at the Hopewell on Wednesday accords with one’s intention (in the sense that it fulfils it), while one’s staying at home to grade logic exams fails to accord with it. Say that one believes that the cat is on the mat. Then, the state of affairs in which the cat is on the mat accords with one’s belief (in the sense that it renders it true), while the state of affairs in which the cat is on the roof does not. Say that one desires to smoke a Bolivar Number 3. Then, one’s smoking a Bolivar Number 3 accords with one’s desire (in the sense that it satisfies it), while one’s smoking a Café Crème does not accord with it. Given this, arguments about rule-following have consequences for our conception of mental content: if an argument shows that there are no facts about what rule an agent is following, it may also show that there are no facts about the contents of a thinker’s mental states.
请注意,在遵循规则的情况下所涉及的协议概念也适用于我们对具有意向内容的心理状态的普遍概念。假设有人打算参加周三在霍普韦尔剧院演出的萨特的戏剧《无出路》。那么,一个人在周三参加霍普韦尔的演出符合他的意图(在实现它的意义上),而他呆在家里改分数逻辑考试则未能符合他的意图。假设一个人相信猫在地毯上。那么,猫在地毯上的情况符合他的信念(在使其成为真实的意义上),而猫在屋顶上的情况则不符合。假设一个人希望抽一支博利瓦 3 号雪茄。那么,一个人抽博利瓦 3 号雪茄符合他的愿望(在满足它的意义上),而抽一支卡菲克里姆则不符合。考虑到这一点,关于遵循规则的论点对我们对心理内容的概念有影响:如果一个论点表明没有关于一个人在遵循什么规则的事实,它也可能表明关于思想者心理状态内容的事实不存在。

Before considering the arguments themselves, we’ll pause to reflect on views about the relative priority of linguistic meaning and mental content, and on what presuppositions are required in order for the arguments to be run.
在考虑这些论点本身之前,我们将暂停一下,反思一下关于语言意义和心理内容相对优先性的观点,以及为了能够运作这些论点需要哪些前提。

In his influential 1989 survey, Paul Boghossian distinguishes between two broad types of view:
在他有影响力的 1989 年调查中,保罗·博戈西安区分了两种广义观点:

  • The Sellarsian View: the notion of linguistic meaning is explanatorily prior to the notion of mental content (Sellars 1956).
    塞拉斯观点:语言意义的概念在解释上优先于心理内容的概念(塞拉斯,1956 年)。
  • The Gricean View: the notion of mental content is explanatorily prior to the notion of linguistic meaning (Grice 1989).
    格莱斯观点:心理内容的概念在解释上优先于语言意义的概念(格莱斯,1989 年)。

Boghossian suggests that, irrespective of which of these views is adopted, it will not be possible to develop a sceptical argument that exclusively targets linguistic meaning. On the Sellarsian View, the conclusion that there are no facts about linguistic meaning will ensure that there are no facts about mental content, since on that view it is from the former sort of fact that the latter sort of fact would have to be derived. On the Gricean view, raising a sceptical doubt about linguistic meaning cannot be done without raising a sceptical doubt about mental content.
博戈西安(Boghossian)认为,无论采用哪种观点,都不可能发展出一个专门针对语言意义的怀疑论证。在塞拉斯(Sellarsian)观点中,关于语言意义没有事实的结论确保了关于心理内容也没有事实的存在,因为在这个观点中,后者的事实必须从前者的事实推导出来。在格赖斯(Gricean)观点中,对语言意义提出怀疑也必然引起对心理内容的怀疑。

We would add a third possible view:
我们将增加第三种可能的观点:

  • The Davidsonian View: the notions of mental content and linguistic meaning are explanatorily interdependent; neither takes explanatory priority over the other (Davidson 1984, 2001).
    戴维森观点:心理内容和语言意义的概念在解释上相互依存;两者之间没有解释优先权(戴维森,1984 年,2001 年)。

Clearly, on the Davidsonian view, one cannot pose a sceptical threat to the existence of facts about the meanings of linguistic expressions without also threatening the existence of facts about mental content, and vice versa. And we can add Boghossian’s further observation:
显然,在戴维森观点看来,一个人不能对语言表达的含义的存在提出怀疑的威胁,而不同时也对心理内容的存在提出怀疑,反之亦然。我们还可以补充博戈西安的进一步观察:

If [the sceptical arguments are] effective at all, they should be as effective against linguistic content as they are against mental content. This is evident from the fact that the arguments construct their skeptical case by exploiting features of content properties, but without exploiting any facts about the putative bearers of those properties. Thus, they would apply to anything said to possess content, whether it was mental or not. (1990 [2008: 62])
如果怀疑论的论证能够生效的话,它们对语言内容的有效性应该与它们对心理内容的有效性一样。这是因为这些论证构建其怀疑案例的方式是通过利用内容属性的特征,而不是利用这些属性的假定承载者的任何事实。因此,它们将适用于任何声称具有内容的东西,无论其是否是心理的。

In what follows, then, we’ll move freely between considering arguments about rule-following, linguistic meaning, and mental content.[2]
在接下来的内容中,我们将自由地在考虑关于遵循规则、语言意义和心理内容的论点之间移动。

2. The Sceptical Argument
2. 怀疑论证

We’ll begin with a brief outline of the argument of Kripke’s sceptic. Suppose that I’ve never dealt with numbers larger than 57.[3] (Given our finite nature and the infinitude of the natural number series, there will always in fact be such a number.) I’m asked to perform the computation 68+57, and I arrive at the answer ‘125’, which I take to be right. However, a “bizarre skeptic” (Kripke 1982: 8) questions my certainty. She suggests that in the past I used ‘plus’ and ‘+’ to mean a different function, which she calls “quaddition”. Quaddition yields the same result as addition if the numbers are lower than 57, and 5 otherwise, so the correct result of the aforementioned computation is ‘5’, not ‘125’. I should answer ‘5’ if I intend to use ‘plus’ in the same way in which I have been using it in the past, or so the sceptic suggests.
我们首先简要概述克里普基的怀疑论。假设我从未处理过大于 57 的数字。 [3] (考虑到我们的有限性和自然数序列的无限性,实际上总会存在这样的数字。)有人要求我进行计算“ 68+57 ”,我得到答案“125”,我认为这是正确的。然而,“奇怪的怀疑论者”(克里普基,1982 年:8)质疑我的确定性。她提出,过去我使用“加”和“+”表示不同的函数,她称之为“四加”。四加如果数字小于 57,则产生与加法相同的结果,否则为 5,因此上述计算的正确结果是“5”,而不是“125”。如果我打算像过去一样使用“加”,那么根据这位怀疑论者的建议,我应该回答“5”。

Kripke allows that the sceptic’s proposal is “absolutely wild”, and that she is “crazy” if she “proposes [her] hypothesis sincerely”. He grants, however, that it is not logically impossible, and so “there must be some fact about my past usage that can be cited to refute” that hypothesis (1982: 9). That is, there must be some fact about my past usage that determines that I meant addition by ‘plus’ in the past, and thus that (again, assuming that I intend to use the expression in the same way I have been using it so far) I should answer ‘125’ rather than ‘5’. Importantly, the sceptic does not question my memory concerning past use; indeed, she goes as far as to allow that the exercise of my cognitive powers is faultless, and that I have access to all the facts about my mind and behaviour that are potentially constitutive of my meaning one thing rather than another (1982: 14). Her thought is that if I am not able, even in such cognitively ideal conditions, to provide the fact in virtue of which I mean addition, a fact that properly singles out the function of addition rather than the function of quaddition, it is because there is no such fact. Furthermore, the focus is on past use because “if I use language at all, I cannot doubt coherently that ‘plus’, as I now use it, denotes plus” (1982: 13). But, if the sceptic’s challenge succeeds, it can be generalized, for “if there was no such thing as my meaning plus rather than quus in the past, neither can there be any such thing in the present” (1982: 21). (The fact that the sceptic grants the idealisation of our cognitive powers in the way that she does shows that her argument is not of a piece with the argument of the epistemological sceptic, who is concerned with whether our actual cognitive capacities can lead to knowledge. See Boghossian 1989 [2002: 150]. For dissent over this point, see Ginsborg 2018. Martin Kusch takes the argument to be metaphysical [2006: xiv], but, in contrast with Boghossian, he takes the dialogic setting to play an essential role in it.)
克里普克承认怀疑论者的提议是“完全荒谬的”,如果她“真诚地提出了[她的]假设”,那她是“疯了”。然而,他承认,这并非逻辑上不可能,因此“必定存在一些关于我过去使用的事实,可以被引用来驳斥”那个假设(1982: 9)。也就是说,关于我过去使用的事实必须确定我在过去用“加法”时的意思,因此(再次假设我打算以与迄今为止相同的方式使用该表达)我应该回答“125”而不是“5”。重要的是,怀疑论者并不质疑我对过去使用的记忆;事实上,她甚至允许我认知能力的运用是无误的,并且我可以获取所有关于我的心智和行为的事实,这些事实可能构成我表示一种而不是另一种意思(1982: 14)。她的想法是,如果在这种认知理想条件下,我不能提供构成我意味着加法而不是四加法的事实,这是因为没有这样的事实。 此外,重点在于过去的使用,因为“如果我使用语言,我就无法怀疑‘加’,就像我现在使用的那样,指的是加”(1982 年:13)。但是,如果怀疑论者的挑战成功,它可以被概括为“如果过去没有我意味着加而不是 quus,那么现在也不可能有这样的事情”(1982 年:21)。(怀疑论者以她所做的方式承认我们的认知能力的理想化,这表明她的论点与认识论怀疑论者的论点不一致,后者关注我们的实际认知能力是否能够导致知识。参见 Boghossian 1989 [2002:150]。对于这一观点的异议,请参见 Ginsborg 2018。Martin Kusch 认为这一论点是形而上学的[2006:xiv],但与 Boghossian 相反,他认为对话环境在其中扮演着重要角色。)

As we’ll see, the search for a fact fails, and the sceptic concludes that “the entire idea of meaning vanishes into thin air” (Kripke 1982: 22). Kripke rejects this paradoxical conclusion as “insane and intolerable” (1982: 60) and “incredible and self-defeating” (1982: 71), and goes on to develop, on behalf of Wittgenstein, a sceptical solution, which he takes to be similar in some respects to David Hume’s solution to the sceptical problem about causation (1982: 4; 62–69), and which purports to conceive of meaning in a way that does not lead to paradox. We discuss the sceptical solution in section 3. (George Wilson was the first to insist on the significance of the distinction between the parts of the sceptic’s position Kripke’s Wittgenstein accepts and the parts he rejects—between the basic sceptical conclusion, according to which there are no facts about a speaker of the sort that the sceptic is seeking that constitute her meaning something by an expression, and the radical sceptical conclusion, according to which no one means anything by an expression [Wilson 1994, 1998]. We will return to Wilson’s view in section 3.3.)
正如我们将看到的那样,对事实的搜索失败了,怀疑论者得出结论:“意义的整个概念消失得无影无踪”(克里普克,1982 年:22)。克里普克拒绝这个悖论性的结论,称其为“疯狂且无法容忍”(1982 年:60),并称其为“难以置信且自我戕害”(1982 年:71),然后继续代表维特根斯坦提出怀疑解决方案,他认为在某些方面类似于大卫·休谟对关于因果关系的怀疑问题的解决方案(1982 年:4;62-69),并试图以一种不会导致悖论的方式来构思意义。我们在第 3 节讨论这种怀疑解决方案。(乔治·威尔逊是第一个坚持怀疑论者立场的各个部分,克里普克的维特根斯坦接受和拒绝部分之间区别重要性的人,即根据怀疑论者所寻找的一种说话者的事实构成她通过表达某事的意义的基本怀疑论者结论和根本怀疑论者结论,根据这些结论,没有人通过表达任何意思(威尔逊,1994 年,1998 年)。我们将在第 3.3 节回顾威尔逊的观点。)

Why does Kripke take the search for candidate meaning facts to fail? Recall that all my previous uses of ‘+’ are compatible with my meaning quaddition, which is what enables the sceptic to say that I meant that all along. One might complain that the challenge operates from a “ridiculous model of instruction” (1982: 15), which fails to take into account that to be a competent adder is to have internalized a general instruction or rule that is now “engraved on my mind as on a slate” and which “justifies and determines my present response” (1982: 15–16). But the sceptic will reply that the worry can be raised again with respect to the general instruction or rule, which is just as susceptible to being interpreted in a deviant way as the initial expression. Kripke then considers a variety of other candidates, which are the kernel of various philosophical theories, and argues, on behalf of the sceptic, that none of them fit the bill. Among the facts considered are my being disposed to produce the sum (1982: 22–32), my instantiating a machine whose operations embody the function of addition (1982: 32–35), the simplicity of the addition hypothesis (1982: 37–39), my having a distinctive experience “with its own special quale, known directly” through introspection (1982: 41), my having an image in mind that supposedly singles out addition (1982: 42), my being in a primitive, irreducible state of meaning addition (1982: 51–53), and my grasping an abstract entity, such as a Fregean sense, which singles out addition (1982: 53–54). None of them, Kripke argues, is successful in ruling out the skeptic’s hypothesis that I meant quaddition rather than addition.
为什么克里普克认为寻找候选含义事实会失败?回想一下,我之前对‘+’的所有使用都与我的含义四加法相容,这正是怀疑论者能够说我一直是这么认为的原因。有人可能会抱怨这种挑战来自于一个“荒谬的指导模式”(1982: 15),它未能考虑到,成为一个称职的加法者意味着已经内化了一条普遍的指导或规则,现在这条规则“象在我的头脑里刻着的在板上一样”,并且“证实并决定了我的当前反应”(1982: 15-16)。但是怀疑论者会回答说,对于普遍的指导或规则,这种担忧可以再次提出,因为它与最初的表达一样容易被解释为偏离的方式。克里普克随后考虑了各种其他候选者,它们是各种哲学理论的核心,并代表怀疑论者主张,这些候选者都不符合要求。 在考虑的事实中,包括我有意生成总和(1982 年:22-32),我实例化一个机器,其操作体现加法函数(1982 年:32-35),加法假设的简单性(1982 年:37-39),我通过内省直接认识的独特经验,“具有自己特殊的质性”(1982 年:41),我心中有一个形象,据称能单独表达加法(1982 年:42),我处于原始、不可化简的意义增加状态(1982 年:51-53),以及我理解一个抽象实体,如弗雷格的意义,用以单独表达加法(1982 年:53-54)。克里普克认为,这些都未能成功排除怀疑论者的假设,即我意思是“四加”而非加法。

As mentioned above, dispositionalism and non-reductionism are the most prominently discussed proposals, and we’ll consider them more carefully in section 4 and section 5, respectively. At this stage, we should ask the following question: what are the conditions that a candidate meaning fact must meet? There is controversy in the literature about their nature and plausibility.
如前所述,性情主义和非简约主义是最广泛讨论的提议,我们将分别在第 4 节和第 5 节更仔细地考虑它们。在这个阶段,我们应该问以下问题:候选意义事实必须满足哪些条件?关于它们的性质和可信性在文献中存在争议。

2.1 The Extensionality Condition
2.1 外延性条件

Here is how Kripke first lays out the two conditions:
这里是克里普克首先阐述的两个条件:

An answer to the skeptic must satisfy two conditions. First, it must give an account of what fact it is (about my mental state) that constitutes my meaning plus, not quus. But further, there is a condition that any putative candidate for such a fact must satisfy. It must, in some sense, show how I am justified in giving the answer ‘125’ to 68+57. The ‘directions’ … that determine what I should do in each instance, must somehow be ‘contained’ in any candidate for the fact as to what I meant. Otherwise, the skeptic has not been answered when he holds that my present response is arbitrary. (1982: 11)
回答怀疑者必须满足两个条件。首先,它必须说明构成我的意义加上,而不是 quus 的事实是什么(关于我的心理状态)。但进一步,任何可能的候选事实都必须满足一个条件。它必须在某种意义上显示我如何有理由对“ 68+57 ”给出答案“125”。在每种情况下确定我应该做什么的“方向”……必须以某种方式“包含”在我所指的事实的任何候选人中。否则,当他认为我的现在回应是任意的时候,怀疑者就没有被回答。(1982 年:11)

To begin with, Kripke claims that whatever fact makes it the case that a speaker means addition by ‘+’ must single out the addition function, as opposed to the quaddition function, as what is meant. It follows from this that the putative meaning-constituting fact must account for the conditions of correct application of ‘+’. In other words, the fact in which my meaning addition by ‘+’ consists must single out ‘4’ as the correct response to 2+2, ‘110’ as the correct response to 55+55, ‘125’ as the correct response to the query 68+57, and so on. Meaning facts render the applications of expressions correct or incorrect, and so a fact cannot count as a meaning-constituting fact unless it does this. This is an uncontroversial claim, which most philosophers accept. As Simon Blackburn notes, this
首先,克里普克声称,使说话者通过‘+’表达加法的事实必须单独确定加法函数,而不是四加函数,作为所指的对象。由此可推出,所谓的构成意义的事实必须解释‘+’的正确应用条件。换句话说,我通过‘+’表示加法的意义所在之事实必须将‘4’单独确定为对‘ 2+2 ’的正确回答,将‘110’单独确定为对‘ 55+55 ’的正确回答,将‘125’单独确定为对‘ 68+57 ’的查询的正确回答,依此类推。意义事实使表达式的应用正确或不正确,因此,除非事实能做到这一点,否则不能将其视为构成意义的事实。这是一个无争议的论断,大多数哲学家都接受。正如西蒙·布莱克本所指出的那样,

distinguishes the production of terms from mere noise, and turns utterance into assertion—into the making of judgement,
区分术语的产生与简单的噪音,并将话语转化为断言—转化为判断的制造

and so  所以

it is not seriously open to a philosopher to deny that, in this minimal sense, there is such a thing as correctness and incorrectness. (1984 [2002: 29]; see also Wikforss 2001: 206; Hattiangadi 2006, 222; Glüer & Wikforss 2009: 35; see Travis 2006 for a dissenting view)
哲学家不太可能否认,在这种最低限度的意义上,存在着正确与不正确之分。(1984 [2002: 29]; 另见 Wikforss 2001: 206; Hattiangadi 2006, 222; Glüer & Wikforss 2009: 35; 另见 Travis 2006 提出的不同观点)

Moreover, this is something on which those who seek to offer a reductive account of meaning (e.g., Fodor 1990; Millikan 1984) and those who are sceptical about the prospects of reduction (e.g., Boghossian 1989; Verheggen 2011; Bridges 2014) seem to agree.
此外,对于那些试图提供关于意义还原的解释(例如,Fodor 1990 年;Millikan 1984 年)以及对于那些对还原前景持怀疑态度的人(例如,Boghossian 1989 年;Verheggen 2011 年;Bridges 2014 年),似乎存在共识。

We’ll call this first condition the extensionality condition. In the case of a predicate like ‘green’, for example, it requires that the fact which constitutes its meaning determines the appropriate class of things to which ‘green’ is correctly applicable. This will be the class of green things as opposed, say, to the class of grue things (where an object is grue if and only if it is green before some specified time t and blue thereafter), and in the arithmetical case on which we have focussed so far, the extension of ‘+’ will contain the triple 57,68,125 (and not the triple 57,68,5).
我们将这个第一个条件称为外延性条件。例如,在像“绿色”这样的谓词的情况下,它要求构成其意义的事实确定了适当的类别,以便正确应用于“绿色”的事物。这将是绿色事物的类,与蓝色事物的类相对立(例如,如果对象在某个特定时间点之前是绿色,之后是蓝色,则该对象是蓝绿色)。在我们迄今为止关注的算术案例中,“+”的外延将包含三元组 57,68,125 (而不是三元组 57,68,5 )。

To see how a candidate meaning-constituting fact might fail the extensionality condition, consider a simple form of dispositional theory of meaning which proposes as constitutive of my meaning addition the fact that I’m disposed to answer with the sum (as opposed, say, to the quum) when faced with arithmetical queries of the form x+y. The sceptic argues that this fact doesn’t single out the addition function:
要看一个候选的构成意义事实如何不符合外延性条件,考虑一种简单的倾向理论,该理论提出构成我意义的添加事实是,当我面对形如“ x+y ”的算术查询时,我倾向于用总和来回答(而不是举例说)。怀疑论者认为,这个事实并不能单独确定出加法函数:

Let ‘quaddition’ be redefined so as to be a function which agrees with addition for all pairs of numbers small enough for me to have any disposition to add them, and let it diverge from addition thereafter (say, it is 5). Then, just as the skeptic previously proposed the hypothesis that I meant quaddition in the old sense, now he proposes quaddition in the new sense. A dispositional account will be impotent to refute him. As before, there are infinitely many candidates the skeptic can propose for the role of quaddition. (Kripke 1982: 27)
让“quaddition”重新定义为一个函数,该函数在对于所有足够小以至于我有可能去加它们的数对上与加法一致,并且在此之后偏离加法(比如,它是 5)。然后,正如怀疑论者先前提出的假设我是指“quaddition”在旧意义上,现在他提出新意义下的“quaddition”。一个处置账户将无力驳斥他。像以前一样,怀疑论者可以无限提出候选者担任“quaddition”的角色。(克里普克 1982: 27)

We’ll consider whether dispositionalism can muster resources to deal plausibly with this problem concerning the satisfaction of the extensionality condition below (in section 4).
我们将考虑位相主义是否能够调动资源来合理处理关于满足下文中外延条件问题的问题(见第 4 节)。

2.2 The Normativity Condition
2.2 规范性条件

According to a prominent line of thought, the notion of correctness involved in the seemingly platitudinous claim that meaningful expressions have conditions of correct application is intrinsically normative. On this reading, meaning facts are normative facts—they not only sort the applications of expressions into correct or incorrect, but also prescribe how expressions ought to be applied. They issue semantic categorical obligations that bind speakers in determinate ways; the justified applications are precisely those that fulfil these semantic obligations. (The kind of normativity at stake is meaning engendered, rather than meaning determining; it is grounded in meaning, rather than grounding meaning. See Glüer and Wikforss 2009 for this helpful distinction.)
根据一种重要的思路,涉及在看似平凡的断言中所涉及的正确性概念,即意义表达具有正确应用条件的主张是内在规范性的。根据这种解读,意义事实是规范事实——它们不仅将表达的应用分类为正确或不正确,还规定了表达应如何应用。它们提出语义范畴义务,以确定方式约束说话者;合理的应用恰恰是那些履行这些语义义务的应用。(所涉及的规范性是由意义引发的,而不是决定意义的;它植根于意义,而不是作为意义的基础。详见 Glüer 和 Wikforss 2009 年对此有益的区分。)

To illustrate how the second condition, thus construed, constrains accounts of meaning, let us again consider Kripke’s discussion of dispositionalism. He thinks that the dispositionalist offers “a descriptive account” of the relation between what one means by an expression and one’s uses of that expression, but that “this is not the proper account of the relation, which is normative, not descriptive” (1982: 37). More generally, Kripke says, “the relation of meaning and intention to future action is normative, not descriptive” (1982: 37). Among the commentators who read the second condition in this way are Wikforss 2001, Glüer and Wikforss 2009, Hattiangadi 2006 and 2007, and Miller 2011, 2012. Thus construed, the sceptical argument can be compared to arguments in metaethics that purport to establish, by drawing on J. L. Mackie’s (1977) argument from queerness concerning the seemingly problematic metaphysical and epistemological status of moral properties, an error-theory of moral judgment (Miller 2010a, 2020). (See the entry on moral anti-realism.)[4]
为了说明如何理解第二条件如何约束对意义的解释,让我们再次考虑克里普克关于倾向主义讨论的例子。他认为,倾向主义者提供了关于一个表达式的含义和一个人对该表达式使用的关系的“描述性解释”,但“这不是关系的恰当解释,这是规范性的,而不是描述性的”(1982: 37)。更普遍地说,克里普克认为,“意义和意图与未来行动的关系是规范性的,而不是描述性的”(1982: 37)。那些将第二条件理解为这种方式的评论者包括 Wikforss 2001 年,Glüer 和 Wikforss 2009 年,Hattiangadi 2006 年和 2007 年,以及 Miller 2011 年和 2012 年。因此,这种理解方式下,怀疑论的论证可以与关于元伦理学的论证进行比较,后者试图通过利用 J. L. 麦基(1977 年)关于道德属性看似问题重重的形而上学和认识论地位的奇异性的论证,来建立道德判断的错误理论(Miller 2010a, 2020)。

Kripke’s discussion has resulted in a vigorous debate about whether meaning really is normative, as well as about how the normativity of meaning is best understood. For a defence of the claim that meaning is normative, see Whiting 2007, 2009, 2016 (note that Whiting focuses on the idea that meaning facts engender permissions to apply words correctly and obligations not to apply them incorrectly, though both the permissions and the obligations are defeasible). For criticism of the view that meaning is normative, see Fodor 1990, Glüer and Pagin 1998, Glüer 1999, Wikforss 2001, Boghossian 2005, Miller 2006, Hattiangadi 2006, 2007, and Glüer and Wikforss 2009. (See also the entry on the normativity of meaning.) Some philosophers seek to carve a middle ground between the normativist and the anti-normativist positions—for instance, by claiming that meaning facts are essentially justificatory (Gampel 1997), or that they have hypothetical implications that are essential to them, thus being fundamentally unlike natural facts, which may be hypothetically normative only accidentally (Verheggen 2011; Chapter 2 of Myers and Verheggen 2016). Hannah Ginsborg proposes a novel conception of normativity as more basic than rules (Ginsborg 2011b, 2012, forthcoming), which we shall briefly discuss in the last section.
克里普基的讨论引发了关于意义是否真正具有规范性的激烈辩论,以及关于如何最好理解意义规范性的辩论。为了辩护意义具有规范性的主张,请参阅 Whiting 2007 年、2009 年、2016 年的作品(注意,Whiting 集中讨论的是意义事实导致正确使用词语的许可和不正确使用词语的义务,尽管这些许可和义务都是可推翻的)。对认为意义具有规范性观点的批评,请参阅 Fodor 1990 年、Glüer 和 Pagin 1998 年、Glüer 1999 年、Wikforss 2001 年、Boghossian 2005 年、Miller 2006 年、Hattiangadi 2006 年、2007 年以及 Glüer 和 Wikforss 2009 年的作品。(另请参阅有关意义规范性的条目。)一些哲学家试图在规范主义者和反规范主义者立场之间开辟中间地带——例如,声称意义事实本质上是证明性的(Gampel 1997 年),或者它们具有假设性的涵义,这对它们来说是基本的区别于自然事实,后者可能仅仅偶然具有假设性(Verheggen 2011 年;Myers 和 Verheggen 2016 年第 2 章)。 汉娜·金斯伯格提出了一种新颖的规范性概念,认为规范性比规则更基本(金斯伯格 2011b,2012,即将出版),我们将在最后一节简要讨论。

Some commentators take the normativity condition to amount to an agential requirement, one that primarily concerns the applications or uses of expressions. The thought is that meaningful uses of expressions are not arbitrary—they are not unjustified leaps in the dark. An adequate conception of meaning must be able to account for this. This view of the normativity condition claims to shed light on Kripke’s numerous appeals to the metaphor of blindness (1982: 10, 15, 17, 23, 87). Kusch thinks that the requirement of non-blindness is best understood as falling under the purview of semantic normativity (which, contra the normativist interpreters, he does not take to involve categorical obligations), and that it is best understood as indicating that the speaker’s “meaning-constituting mental state guides and instructs” her on how to apply the expression, that the speaker “can refer to this mental state in order to justify her use” of the expression, and that this state
有些评论者认为,规范性条件等同于一种行动要求,主要涉及表达的应用或使用。这种观点认为,表达的有意义使用并非任意的——它们不是在黑暗中不经过理由的跳跃。对意义的充分概念必须能够解释这一点。这种对规范性条件的观点声称能够阐明克里普克对“盲目”隐喻的众多引述(1982 年:10、15、17、23、87)。库施认为,非盲目的要求最好理解为语义规范性的范畴(与规范主义解释者相反,他认为这并不涉及范畴义务),最好理解为指示说话者的“意义构成心态引导和指导”她如何应用表达,说话者“可以参考这种心态来证明她使用表达的理由”,并且这种状态

not only justifies certain applications—in the sense that meaning addition justifies ‘125’ in answer to 68+57=?—it also justifies the way in which the answer is usually produced. (Kusch 2006: 8–9, italics added)
不仅仅是在某些情况下证明合理——在这种意义上,意义的增加证明了对“ 68+57=? ”的回答中的“125”——它还证明了通常产生答案的方式。(Kusch 2006: 8–9,斜体添加)

Along similar lines, it has been suggested that the meaning-constituting facts be able to accommodate the idea that “the meaningful use of words must be revealed as intentional” (Sultanescu and Verheggen 2019, 13; Sultanescu forthcoming). The paradox has also been interpreted as belonging
在类似的思路下,有人建议认为构成含义的事实应能够适应“必须将单词的有意义使用表现为有意图”这一观念(Sultanescu and Verheggen 2019, 13; Sultanescu forthcoming)。这个悖论也被解释为属于

to the philosophy of rational explanation, of explanations that account for what people do or think by citing their reasons for doing or thinking so. (Bridges 2014: 249; see also Bridges 2016)
对理性解释的哲学,即通过引用人们做某事或思考某事的理由来解释他们的行为或思维。(Bridges 2014: 249; 另见 Bridges 2016)

Some interpreters take the sceptical argument to involve the imposition of an epistemological constraint—a constraint related to the epistemic justification of semantic judgments, rather than to the rationality of the applications of expressions. Warren Goldfarb notes that Kripke “does seem to mean that the justifications must in some sense be transparent” (1985 [2002: 98]). José Zalabardo takes Kripke to be demanding that the meaning-constituting facts provide speakers with justification for their judgments about the correctness of the applications of their predicates. Justification is construed in an internalist sense: speakers possess the relevant justification if the procedure through which they decide whether predicates apply to objects involved “conscious engagement with the facts that determine how these questions should be answered” (1997 [2002: 286]). (See also Jackman 2003, Guardo 2012, and Merino-Rajme 2015). Crispin Wright proposes a more specific epistemic constraint, namely, that of accounting for the seemingly puzzling fact that the epistemology of meaning is first-person authoritative even though meaning something by an expression is in crucial respects akin to having a dispositional trait. In Wright’s words, the constraint is to explain
有些解释者认为怀疑论的论证涉及对认识论约束的施加——这种约束与语义判断的认知正当性有关,而不是与表达式应用的合理性有关。沃伦·戈德法布指出,克里普克似乎“确实意味着这些正当化必须在某种意义上是透明的”(1985 年[2002: 98])。何塞·萨拉巴尔多认为克里普克要求构成含义的事实为发言者提供对其关于谓词应用正确性的判断的正当化。正当化被解释为内在主义意义上的:如果发言者在决定是否将谓词应用于对象时的过程中“有意识地参与了决定如何回答这些问题的事实”,则发言者拥有相关的正当化(1997 年[2002: 286])。(另见杰克曼 2003 年,瓜尔多 2012 年和梅里诺-拉赫梅 2015 年。) 克里斯平·赖特提出了一个更具体的认识论限制,即要解释一个看似令人困惑的事实,即意义论的认识论是第一人称权威的,尽管通过表达意思在关键方面类似于具有一种倾向性特征。莱特说,限制在于解释

how it is possible to be effortlessly, non-inferentially and generally reliably authoritative about psychological states which have no distinctive occurrent phenomenology and which have to answer, after the fashion of dispositions, to what one says and does in situations so far unconsidered. (2001: 150)
如何能够毫不费力、非推理地并且通常可靠地对没有显著当前现象学的心理状态拥有权威性,这些状态必须像倾向性一样回答在尚未考虑的情境中所说和所做的事情。

Note that all the construals of the second condition appear to put pressure on dispositionalism over and above that exerted by the extensionality condition. Prima facie, dispositional facts are facts about what we will or would do, not about what we ought to do; dispositional facts appear not to be essentially justificatory or hypothetically prescriptive; dispositions do not justify or rationalise their manifestations, and in making semantic judgements we do not typically engage with facts about our linguistic dispositions; lastly, it is unclear how a dispositional account of meaning could be rendered consistent with its intuitive first-person epistemology. See section 4 for further discussion of dispositionalism. 重试    错误原因

3. The Sceptical Solution
3. 怀疑主义解决方案

If the sceptical argument is cogent, it seems to follow that there are no meaning-constituting facts, no facts in virtue of which linguistic expressions mean one thing rather than another. As noted above, this appears to imply the paradoxical conclusion that “the entire idea of meaning vanishes into thin air” (1982: 22). Kripke distinguishes between two broad ways in which one might attempt to avoid this conclusion (1982: 66–7). On the one hand, one might provide a straight response, by identifying some meaning-constituting fact of the sort called into question by the sceptic. The various proposals discussed briefly in section 2 above are instances of straight responses. The two most prominent types of straight response in the literature—reductive dispositionalism and non-reductionism—are discussed in more detail in section 4 and section 5 below. On the other hand, one might provide a sceptical response. That is, one might concede that there are no meaning-constituting facts of the sort demanded by the sceptic but deny that this leads to a paradoxical conclusion. In this section, we shall focus on this strategy.
如果怀疑论的论证是有说服力的,似乎可以得出这样的结论:没有构成含义的事实,也没有事实支持语言表达的含义与其他含义不同。如上所述,这似乎意味着一个悖论性的结论:“整个含义的概念化作了泡影”(1982 年:22)。克里普克区分了两种可以尝试避免这一结论的广泛方式(1982 年:66-7)。一方面,可以提出一个直接的回应,通过确定一些与怀疑者质疑的那种构成含义的事实。上述第 2 节简要讨论的各种提案是直接回应的实例。文献中两种最突出的直接回应类型——归纳式处置主义和非归纳式——将在下文的第 4 节和第 5 节中更详细地讨论。另一方面,可以提出怀疑性的回应。也就是说,可以承认没有构成含义的事实,这种事实是怀疑者所要求的,但否认这导致了一个悖论性的结论。在本节中,我们将专注于这种策略。

The proponent of the sceptical solution can be understood as rejecting eliminativism about our practices of ascribing meaning. Mirroring parallel discussions in metaethics, the two most obvious paths available to her involve providing either an error-theoretic account or a non-factualist account of ascriptions of meaning. We shall follow Boghossian in viewing these paths as forms of irrealism about meaning, content and rules (Boghossian 1989, 1990). A prominent general line of argument in the recent literature suggests that irrealist views of any area make presuppositions that irrealist views of meaning and content are bound to deny, so that irrealism about meaning and content is ultimately incoherent (Boghossian 1989, 1990; Hattiangadi 2007, 2017, 2018; Miller 2011, 2015a, 2020). We’ll illustrate this general line of attack by sketching an argument to the effect that, regardless of whether one pursues an error-theoretic or a non-factualist approach, adopting irrealism leads inexorably to an “insane and intolerable” and “incredible and self-defeating” form of eliminativism on which the notions of meaning and content do turn out to “vanish into thin air”. We’ll then briefly consider an alternative way of providing a sceptical response, which aims to revise the conception of meaning fact that is at work in the sceptic’s mindset.
怀疑解决方案的提倡者可以被理解为拒绝消除我们赋予意义的实践。与形而上伦理学中的平行讨论相呼应,她可以选择提供错误理论解释或非事实主义解释赋予意义的方式中的任何一种。我们将跟随博格西安的观点,将这些方式视为对意义、内容和规则的非现实主义形式。最近文献中突出的一种一般性论证线索表明,对任何领域的非现实主义观点都会假设非现实主义观点关于意义和内容的前提,而非现实主义关于意义和内容的观点最终是不连贯的。我们将通过简要勾画的论证来说明这种一般性攻击线索,不论是追求错误理论还是非事实主义方法,采用非现实主义都不可避免地导致“疯狂和无法忍受”的“难以置信和自相矛盾”的消除主义形式,在这种形式中,意义和内容的概念最终“变得无影无踪”。 我们接下来会简要考虑一种提供怀疑回应的替代方式,旨在修订怀疑者心态中运作的意义事实观念。

3.1 Error Theories 3.1 错误理论

Suppose we adopt an error theory: the view that all atomic, positive statements ascribing meaning, content, or the following of a rule, are false. While some error theories are eliminativist (e.g., Churchland 1981 on propositional attitudes), the error theorist need not subscribe to eliminativism. For instance, J.L. Mackie (1977) argues that although moral judgements are uniformly false, eliminativism can be avoided given that some moral judgements are such that their acceptance facilitates securing the benefits of social cooperation in circumstances where “the limitation of men’s sympathies” (1977: 108) threaten their attainment.[5] On this view, even though our practice of making moral judgments results in falsehoods, it meets a subsidiary norm (or norms) in terms of whose satisfaction its pragmatic utility can be secured.[6] Might an error theorist about meaning, content and rule-following attempt to avoid eliminativism by following a similar strategy?
假设我们采取了一个错误理论:即所有关于意义、内容或遵循规则的原子正面陈述都是错误的观点。虽然一些错误理论是排除论的(例如,某些关于命题态度的错误理论,如 Churchland 1981),但错误理论者不一定要拥护排除论。例如,J.L. Mackie(1977)认为,尽管道德判断普遍是错误的,但可以避免排除论,因为有些道德判断是这样的,即它们的接受有助于在“限制人们的同情心”的情况下确保获得社会合作的好处(1977 年:108)。根据这一观点,尽管我们作出道德判断的做法导致了错误,但它符合一个辅助规范(或多个规范),通过满足这些规范,可以确保其实用性。对于意义、内容和规则遵循的错误理论者,是否可以通过采用类似的策略来避免排除论?

It might seem that there is room for this approach. In the case of meaning, the subsidiary norm might be something like the following: one ought to assert “Jones means addition by ‘+’” only when Jones’s
这种方法似乎是有空间的。在意义的情况下,辅助规范可能是这样的:只有当琼斯的

particular responses to arithmetical queries agree with those of the community in enough cases, especially the simple ones (and if his “wrong” answers are not often bizarrely wrong, as is ‘5’ for 68+57, but seem to agree with ours in procedure, even when he makes a “computational mistake”). (Kripke 1982: 92)
对算术问题的特定回答在足够数量的情况下与社区的回答保持一致,特别是简单的问题(即使他的“错误”答案不常常是荒谬地错误,如将‘5’误写为‘ 68+57 ’,而是在过程上与我们的回答一致,即使他在计算上出了“计算错误”)。(克里普克,1982 年: 92)

This norm would thus be cashed out in terms of agreement with respect to inclinations “to go on” in certain ways, and the utility of our complying with it would be that it enables us to make helpful discriminations—for example, when seeking to buy five apples—between grocers whose inclinations match ours and grocers with “bizarre” quus-like inclinations.[7]
因此,这种规范可以通过倾向性“继续”某些方式的一致达成,我们遵守它的效用在于它使我们能够进行有益的区分——例如,在寻求购买五个苹果时——能够区分出与我们倾向相符的杂货店和具有“奇特”quus 样倾向的杂货店。

While this strategy might seem to some to be promising in the moral case, it faces special problems in the case of meaning, content and rules (Boghossian 1989, 1990; Miller 2015a). In order for the strategy to be able so much as to be pursued, there has to be such a thing as complying or failing to comply with a subsidiary norm—and so, a fortiori, such a thing as complying or failing to comply with a norm as such. An error theorist about rule-following, however, denies precisely that there are facts about norm-compliance and non-compliance. Having argued that all statements about rule-compliance are false, the error theorist apparently lacks the resources for telling a story about the pragmatic utility of our continuing to engage in the practice of making judgements about rule-following. The upshot seems to be that, if we accept that there are no facts about speakers in virtue of which expressions have meaning, embracing an error theory will not prevent the notion of meaning from “vanishing into thin air”.
尽管这种策略在道德案例中似乎有所希望,但在涉及意义、内容和规则的情况下,它面临着特殊问题(Boghossian 1989, 1990; Miller 2015a)。为了使这种策略能够被追求,必须存在遵守或未遵守辅助规范的事物,因此,更进一步地,存在遵守或未遵守规范本身的事物。然而,关于遵循规则的错误理论家恰恰否认存在关于规范遵从和不遵从的事实。在论证了关于规则遵从的所有陈述都是错误的之后,错误理论家似乎缺乏讲述我们继续参与对规则遵循进行判断实践的实用性的资源。总的来说,如果我们接受在讲话者身上不存在关于表达具有意义的事实,那么接受错误理论将无法阻止意义的概念“消失于空气中”。

3.2 Non-Factualist Theories
3.2 非事实主义理论

One might attempt to avoid eliminativism about meaning by embracing a different type of irrealism, namely, non-factualism. A non-factualist about a domain maintains that judgments and claims made within that domain are not in the business of stating facts. Indeed, the standard view in the secondary literature is that Kripke’s Wittgenstein himself is proposing a form of semantic non-factualism in the sceptical solution outlined in chapter 3 of Kripke (1982). See, e.g., McGinn (1984), Wright (1984), Boghossian (1989), and Hale (2017). Might semantic non-factualism afford us a way to embrace the conclusion of the sceptical argument while avoiding eliminativism?
有人或许试图通过接受一种不同类型的非现实主义,即非事实主义,来避免关于意义的消除论。关于某一领域的非事实主义者认为,在该领域内所作的判断和主张并非在陈述事实。事实上,次级文献中的标准观点是,克里普克的维特根斯坦本人在克里普克(1982)第 3 章中概述的怀疑解决方案中,提出了一种语义非事实主义的形式。参见,例如,麦金(1984)、赖特(1984)、博格西安(1989)和哈尔(2017)。语义非事实主义也许能让我们在接受怀疑论论证的结论的同时避免消除论?

This move faces difficulties parallel to those faced by error theories (Boghossian 1989, 1990; Hattiangadi 2007: chapter 4, 2018; Miller 2011, 2020). The non-factualist about meaning proposes that we construe ascriptions of meaning as having a purpose or function different from that of stating facts. But, insofar as a sentence is regarded as having a function, there is an intelligible distinction between correct and incorrect uses of that sentence; in other words, the sentence is rule-governed. That the notion of correctness in these cases cannot be identified with truth or warranted assertibility makes no difference to the applicability of a generalised version of the extensionality condition, which we outlined in section 2. Thus, suppose that S is a sentence that has a non-descriptive semantic function. Consider these two specifications of the conditions in which uttering S is correct, where R1 and R2 would confer different correctness-values on utterances of S:
这种做法面临的困难与错误理论所面临的困难类似(Boghossian 1989, 1990; Hattiangadi 2007: 第 4 章, 2018; Miller 2011, 2020)。关于含义的非事实主义者建议,我们把对含义的归属理解为具有与陈述事实不同的目的或功能。但是,只要把一个句子视为具有某种功能,就能够明确正确使用和错误使用该句子之间的区别;换句话说,该句子受规则约束。在这些情况下,正确性的概念不能被视为真实或有保证的断言,这对我们在第 2 节中概述的外延性条件的推广版本的适用性没有影响。因此,假设‘ S ’是一个具有非描述性语义功能的句子。考虑以下两种情况的具体规定,在这些情况下发出‘ S ’是正确的,其中 R1R2 会使‘ S ’的发声获得不同的正确性价值:

(a)
Uttering S is correct iff the conditions in which it is uttered accord with R1;
发出“ S ”是正确的,当它被说出的条件符合 R1 时;
(b)
Uttering S is correct iff the conditions in which it is uttered accord with R2.
发出“ S ”是正确的,当且仅当它发出的条件符合 R2

Whatever makes it the case that (a) provides the correctness conditions for utterances of S must rule out (b) as providing those conditions.
无论是什么使得对“ S ”的话语具有正确性条件,必须排除(b)作为提供这些条件的情况。

Now, let’s consider the case of plans, which are expressed in sentences that wear their non-descriptive semantic function on their sleeves. (We choose the case of plans deliberately here, as a view of this kind is mooted in Gibbard 2012.) Take the sentence ‘Let’s write a fugue!’. Consider two possible correctness conditions for it:
现在,让我们考虑计划的情况,这些计划以明显的非描述性语义功能表达在句子中。(我们在这里故意选择计划的情况,因为这种观点在 Gibbard 2012 中被提出。)以句子“让我们写一首赋格!”为例。考虑其可能的两种正确性条件:

(a*) (a*) (a*)
‘Let’s write a fugue!’, as uttered by Glenn, is correct iff Glenn plans to write a fugue
让我们写一首赋格!格伦所说的是正确的,当且仅当格伦计划写一首赋格

and

(b*)
‘Let’s write a fugue!’, as uttered by Glenn, is correct iff either (a) it is time t earlier than t* and Glenn plans to write a fugue or (b) it is time t later than or equal to t* and Glenn plans to write a novel.
“让我们写一首赋格!”,当格伦说出这句话时,它是正确的,当且仅当(a)时间 t 早于 t*,而格伦计划写一首赋格,或者(b)时间 t 晚于或等于 t*,而格伦计划写一部小说。

Glenn’s finite nature together with the infinitude of the temporal sequence ensures that the sceptic will always be able to argue that no fact about Glenn is capable of ruling in something like (a*) and ruling out something like (b*) as the relevant correctness condition. The non-factualist is entitled to regard “Let’s write a fugue!” as having a determinate meaning only if she can provide some fact about Glenn or his speech community that determines that that sentence is governed by (a*) rather than (b*). The same argument can be pressed against the suggestion that “Jones means addition by ‘+’” should be regarded as having some non-descriptive semantic function. The non-factualist will be able to regard “Jones means addition by ‘+’” as having a determinate meaning only if she can provide some fact about Jones or his speech community that determines that that sentence is governed by one rule or correctness condition rather than another.
格伦的有限性与时间序列的无限性确保了怀疑论者始终能够主张,关于格伦的任何事实都无法确定像(a*)这样的某种事实,并排除像(b*)这样的其他相关正确性条件。非事实主义者有权将“让我们写一首赋格曲!”视为具有确定含义,前提是她能够提供关于格伦或其言语社区的某些事实,这些事实决定了该句子受(a*)而非(b*)的控制。同样的论据也适用于“琼斯用‘+’号表示加法”的建议,应被视为具有某种非描述性语义功能的提议。非事实主义者只有在能够提供关于琼斯或其言语社区的某些事实,决定该句子受一种规则或正确性条件的控制而非另一种时,才能将“琼斯用‘+’号表示加法”视为具有确定含义。

So, a non-factualist account of any region of thought and talk, which is committed to the claim that there are no facts of the relevant sort, would seem to presupposes a realist account of meaning, content and rules according to which there are semantic facts, intentional facts, and facts about normative accord. So, a non-factualist account of meaning, content and rules, which is committed to the claim that there are no semantic facts, intentional facts, or facts about normative accord that our semantic, intentional, or normative discourse purports to capture, presupposes a realist account of meaning, content and rules according to which there are semantic facts, intentional facts, and facts about normative accord. It thus faces a charge of incoherence.
因此,任何对思维和言论领域的非事实主义描述,即承认相关类型的事实不存在的说法,似乎预设了一个关于意义、内容和规则的现实主义描述,根据这种描述,存在语义事实、意图事实以及关于规范符合的事实。因此,任何对意义、内容和规则的非事实主义描述,即坚持认为我们的语义、意图或规范性论述没有捕捉到语义事实、意图事实或关于规范符合的事实的说法,预设了一个关于意义、内容和规则的现实主义描述,根据这种描述,存在语义事实、意图事实以及关于规范符合的事实。因此,它面临不一致的指责。

As noted above, we chose “plans” as our stalking horse here in order to make explicit the problem this poses for Allan Gibbard’s (2012) account of meaning (it makes no difference to the argument whether “plans” are taken to be linguistic items possessing meaning or mental states possessing intentional content). A Gibbard-style expressivist approach to an area such as morality denies that there are moral facts, but presupposes that there are facts about meaning, content and normative accord (facts that determine what accords and fails to accord with a plan). A Gibbard-style approach to meaning, content and rules thus, on the one hand, denies that there are semantic facts, intentional facts and facts about normative accord (facts that determine what accords and fails to accord with a plan), and, on the other, presupposes that there are such facts. It thus faces a charge of incoherence.
正如前文所述,我们选择在这里将“计划”作为我们的引子,以明确指出这对艾伦·吉巴德(2012 年)关于意义的论述所构成的问题(无论将“计划”视为具有意义的语言项目还是具有意向内容的心理状态,对论证都没有影响)。吉巴德式的表达主义方法,例如在道德领域,否认存在道德事实,但预设存在关于意义、内容和规范一致性的事实(决定何为符合计划及何不符合计划的事实)。吉巴德式的意义、内容和规则方法,一方面否认存在语义事实、意向事实以及规范一致性的事实(决定何为符合计划及何不符合计划的事实),另一方面却预设这些事实的存在。因此,它面临着不一致性的指控。

A similar objection to Gibbard’s view is outlined in Hattiangadi (2018). According to Hattiangadi, a Gibbard-style expressivist account of moral claims, for example, aims to give an “oblique” explanation of them in terms of the states of mind they express (rather than a “straight” explanation in terms of (moral) states of affairs which potentially render them true). But such an oblique explanation in the moral case presupposes a straight explanation of intentionality. In parallel, an oblique explanation of meaning and intentionality would presuppose a straight explanation of meaning and intentionality, again threatening the view with incoherence. In a reply to Hattiangadi, Gibbard reflects on the strategy of the metaethical expressivist and attempts to use it to counter Hattiangadi’s worry. He writes:
Gibbard 观点的类似反对意见在 Hattiangadi(2018)中概述。根据 Hattiangadi 的说法,例如,Gibbard 式的表达主义对道德主张采取“间接”解释,旨在通过它们表达的心态来解释它们(而不是通过(道德)事实的“直接”解释,这可能使它们成为真实)。但是,在道德案例中进行这样的间接解释,就预设了意向性的直接解释。同样地,对意义和意向性的间接解释将预设意义和意向性的直接解释,再次使该观点面临不连贯的问题。在对 Hattiangadi 的回复中,Gibbard 反思了元伦理表达主义者的策略,并试图利用它来反驳 Hattiangadi 的担忧。他写道:

A parallel can be found in ethics: Suppose we claim that being good consists in being pleasurable. The concept of being pleasurable can be completely non-ethical and naturalistic, but the claim “Being good consists in being pleasurable” is ethical—and so, if Moore and others are right, nonnaturalistic. Is there, then, “a straight or substantive explanation of intentionality”? The correct answer will parallel that for the question, “Is there a straight substantive explanation of being good?” If ethical hedonists are right and being good consists in being pleasurable, then there’s a straight, substantive explanation of being good in the sense of a naturalistic explanation of the property that being good consists in. But the claim “Being good is being pleasurable” isn’t itself naturalistic. If Ayer was right, it amounts to “Hurrah for all and only what”s pleasurable. (Gibbard 2018: 770)
道德伦理学中可以找到一个类比:假设我们声称善良在于令人愉悦。令人愉悦的概念可以完全是非伦理和自然主义的,但是“善良在于令人愉悦”的主张是伦理的——因此,如果摩尔和其他人是对的,那就是非自然主义的。那么,“关于意向性有一个直接或实质性的解释”吗?正确答案将类似于“关于善良有一个直接实质性的解释”的问题。如果伦理享乐主义者是对的,善良在于令人愉悦,那么在自然主义解释善良所包含的属性的意义上,确实有一个直接、实质性的善良解释。但是主张“善良即令人愉悦”本身并非自然主义的。如果艾尔是对的,这相当于“为所有仅为令人愉悦的事情欢呼”。(吉巴德 2018: 770)

We leave assessment of Gibbard’s reply as an exercise for readers.[8]
我们将吉巴德的回复的评估留给读者作为一个练习。 [8]

3.3 An Alternate Form of Factualism
3.3 事实主义的另一种形式

Wilson takes the lesson of the sceptical argument to be not that there are no meaning facts, but rather that a certain conception of such facts, which he calls classical realism, is hopeless, and conceives of the sceptical solution as accommodating meaning facts when conceived in a different way (Wilson 1994; see also Wright 1992: chapter 6). Classical realism is sometimes referred to as semantic platonism, the view that “the meanings of our words are guaranteed by the pre-existing structure of reality” (Pears 1988: 363; cf. Child 2001, Verheggen 2003, Zalabardo 2003, Hanks 2017). What is essential to classical realism or semantic platonism is that the properties that guarantee meaningfulness must be antecedently singled out by individuals (or communities) in order to endow their words with semantic standards (Wilson 1994 [2002: 251]). As Zalabardo puts it, what is required is
威尔逊认为怀疑论论证的教训并非是说没有意义事实,而是某种对这些事实的概念——他称之为经典现实主义,是无望的,并将怀疑论的解决方案构想为以不同方式理解时容纳意义事实(Wilson 1994;参见也 Wright 1992 年第 6 章)。经典现实主义有时被称为语义柏拉图主义,即“我们词语的意义是由现实的预先存在结构保证的”(Pears 1988 年:363;参见 Child 2001 年、Verheggen 2003 年、Zalabardo 2003 年、Hanks 2017 年)。经典现实主义或语义柏拉图主义的本质是必须先由个体(或社群)事先确定保证意义的属性,以赋予其词语语义标准(Wilson 1994 年[2002 年:251])。正如扎拉巴尔多所言,所需的是

a conscious act in which I decide to pair the predicate with the property in such a way that the satisfaction conditions of the predicate, as I mean it, are determined by the instantiation conditions of the property. (2003: 314)
一个自觉的行为,在这个行为中,我决定以这样的方式将谓词与属性配对,以便谓词的满足条件(如我所理解的)由属性的实例化条件确定。(2003 年:314)

Wilson takes the sceptical challenge to reveal that no sense can be made of this idea, for it is not possible for an individual (or community) non-linguistically to single out properties as “the de re subject of her meaning-constituting intentions” (1998: 105). But, to repeat, what this allegedly shows is not that there are no meaning facts, but rather that we must reform our conception of them.
威尔逊接受怀疑的挑战,揭示了这个观念是毫无意义的,因为一个个体(或社区)在非语言上不可能单独确定属性作为“她意义构成意图的主题”(1998 年:105)。但是,需要重申的是,这据称显示的并不是意义事实不存在,而是我们必须改革对它们的理解。

The alternative picture of meaning that Wilson fleshes out conceives of expressions as not connected to properties that serve as “pre-established” standards of correctness” (2003: 181–182), and suggests that “what we mean by [an expression] is something that gets settled only over the course of time” (2003: 186). In response to Wilson’s proposal, it has been argued that it is susceptible to collapsing into a form of subjectivism (Kremer 2000), and that it is untenable, for it falls prey to the sceptical challenge that it purports to bypass (Miller 2010b). It has also been argued, contra Wilson, that classical realism is merely an instance of a more general conception of meaning that takes standards of correctness to be determined by entities—whether abstract properties or real features of the world around us—that are considered independently of how we might describe them linguistically; it is this conception that must be rejected, as it is ultimately responsible for generating the paradox (Verheggen 2003).[9]
Wilson 所描绘的意义的另一种图景将表达看作与作为“预先建立的”正确标准的属性无关,并暗示“我们所指的[表达]是随着时间推移才确定的东西”。对于 Wilson 的提议,有人争辩说它容易崩溃成为一种主观主义形式,并且它是站不住脚的,因为它会受到怀疑挑战的影响,即它声称要绕过的挑战。同时也有人认为,与 Wilson 相反,古典现实主义仅仅是一个更一般的意义观念的一个例证,这一观念认为正确的标准是由被认为是独立于我们如何用语言描述它们的实体确定的——无论是抽象属性还是我们周围世界的真实特征;正是这种观念必须被拒绝,因为它最终是产生这个悖论的原因。

4. Reductive Dispositionalism
4. 减少性倾向主义

The most widely discussed attempt at a straight solution to the sceptical challenge is reductive dispositionalism. According to a simple version of reductive dispositionalism, the fact that Jones has the concept of addition rather than of quaddition is to be identified with (or is constituted by) his disposition to produce the result of adding (and not quadding) the numbers x and y in response to arithmetical queries of the form x+y=?, and the fact that he means cat by ‘cat’ is to be identified with (or is constituted by) his disposition to apply ‘cat’ to cats. (See Horwich 1998, 2010, 2012 for a systematic development of dispositionalism; an answer to Kripke’s challenge is articulated in Horwich 2015.)
最广泛讨论的试图直接应对怀疑挑战的方法是归约倾向主义。根据简单的归约倾向主义版本,琼斯拥有加法概念而非四加概念的事实,被认同为(或者构成为)他在算术查询形式为“ x+y=? ”的问题中,对于数值 x 和 y 相加(而非四加)的结果产生反应的倾向;而他意思“猫”是指他倾向于将“猫”应用于猫的事实。 (关于倾向主义的系统发展,详见 Horwich 1998 年、2010 年、2012 年;对克里普克挑战的回答见 Horwich 2015 年。)

As Boghossian notes (1989 [2002: 164–165]), the general form of dispositionalism targeted by the sceptic covers both conceptual role theories and causal/informational theories. In both cases, the account is intended to be reductive, insofar as the content-determining dispositions are to be characterized in wholly non-semantic and non-intentional terms. The sceptic’s attack on reductive dispositionalist theories is thus an attack on two of the most popular accounts of the determination of content in contemporary philosophy of mind and language.
如博格西安(1989 年 [2002 年:164–165 页])所指出的,怀疑论者针对的一般倾向主义形式涵盖了概念角色理论和因果/信息理论。在这两种情况下,这种解释旨在是还原的,因为决定内容的倾向应当用完全非语义和非意向性的术语来表述。怀疑论者对还原倾向主义理论的攻击因此是对当代心灵哲学和语言学中两种最流行的内容确定理论的攻击。

The sceptic argues that dispositionalist theories face three problems. The first problem—the finitude problem—is that there is a sense in which, much like the totality of our previous linguistic behaviour, our dispositions are finite. Given that the extension of the addition function is infinite, containing a denumerably infinite number of triples x,y,z such that x plus y is identical to z, Jones’s meaning addition by ‘+’ cannot be identified with her dispositions to respond to arithmetical queries since it is simply false that she is disposed to answer with the sum when faced with the query x+y=?. In some (indeed, most) cases, the numbers involved will be so large that Jones’s brain’s capacity for computation is far exceeded, and Jones may even die long before she is able to grasp the relevant numbers. We might follow Boghossian in dubbing such numbers “inaccessible” (2015: 335), and we might define quaddition to be a function that diverges from addition over only inaccessible numbers. In this case, the problem is that, given Jones’s dispositions, it is indeterminate whether he means plus or quus.
怀疑论者认为,倾向论理论面临三个问题。第一个问题——有限性问题——是这样一种情况,就像我们先前的语言行为的总体一样,我们的倾向是有限的。鉴于加法函数的延伸是无限的,包含着一个可数无限数量的三元组 x,y,z ,使得 x 加 y 等于 z,琼斯用‘+’表示的加法意义不能被等同于她对算术问题的反应倾向,因为当面对问题‘ x+y=? ’时,她倾向于回答总和是错误的。在一些(事实上,大多数)情况下,涉及的数字将会如此之大,以至于琼斯的大脑计算能力远远超出,琼斯甚至可能在能够理解相关数字之前就去世了。我们可以像博戈西安(Boghossian)那样称这些数字为“不可及的”(2015: 335),我们可以定义 quaddition 为一种仅在不可及数字上与加法有所不同的函数。在这种情况下,问题在于,鉴于琼斯的倾向,他是指加法还是 quus 是不确定的。

The second problem—the error problem—is that someone might be systematically disposed to make mistakes. Take Smith, who is systematically disposed to miscarry when responding to x+y=? queries. When Smith produces ‘28’ in response to 19+19=?, we want to be able to say that her answer is incorrect in the light of what she means by ‘+’. However, if what she means is determined by her dispositions, we are forced to say that she actually means some non-standard function (one that corresponds to addition with the carrying operation removed), so that her answer ‘28’ is correct.
第二个问题——错误问题——是某人可能有系统性倾向于犯错。比如史密斯,在回答“ x+y=? ”的查询时有系统性倾向于错误。当史密斯在回答“ 19+19=? ”时产生了“28”,我们希望能够说,根据她对“+”的意思,她的答案是不正确的。然而,如果她的意思由她的倾向决定,我们被迫说她实际上是指一些非标准的函数(即去除进位操作的加法),这样她的答案“28”就是正确的。

The third problem—the normativity problem—is that the dispositionalist view seems unable to capture the normativity of meaning. Given what she means by ‘+’, Jones ought to respond to arithmetical queries of the form x+y=? by producing the sum of x and y, but the meaning-constituting fact proposed by the dispositionalist is at most a fact about how she would respond to queries of the relevant form:
第三个问题——规范性问题——是指倾向性观点似乎无法捕捉意义的规范性。根据她对“+”的意思,琼斯应该对形如“ x+y=? ”的算术查询做出响应,产生 x 和 y 的总和,但倾向性主义者提出的构成意义的事实最多只是关于她如何对相关形式的查询做出响应的事实。

Suppose I do mean addition by ‘+’. What is the relation of this supposition to the question how I will respond to the problem 68+57’? The dispositionalist gives a descriptive account of this relation: if ‘+’ meant addition, then I will answer ‘125’. But this is not the proper account of the relation, which is normative, not descriptive. The point is not that, if I meant addition by ‘+’, I will answer ‘125’, but that, if I intend to accord with my past meaning of ‘+’, I should answer ‘125’. Computational error, finiteness of my capacity, and other disturbing factors may lead me not to be disposed to respond as I should, but if so, I have not acted in accordance with my intentions. The relation of meaning and intention to future action is normative, not descriptive. (Kripke 1982: 37)
假设我用‘+’表示加法。这一假设与我如何回应问题‘ 68+57 ’的关系是什么?支持性理论者对这种关系给出了描述性说明:如果‘+’表示加法,那么我会回答‘125’。但这并不是关系的适当说明,这种关系是规范性的,而非描述性的。关键不在于如果我用‘+’表示加法,我会回答‘125’,而在于如果我打算符合我对‘+’的过去理解,我应该回答‘125’。计算错误、我的能力有限以及其他干扰因素可能导致我不能按照应有的方式回应,但如果如此,我并未按照我的意图行事。意义和意图与未来行动的关系是规范性的,而非描述性的。(克里普克 1982: 37)

In section 2, we outlined a number of ways in which a normativity condition might be thought to impose a constraint on accounts of meaning. We suggested that in all of these ways this condition puts at least prima facie pressure on dispositionalist theories of meaning. We will limit ourselves in the remainder of this section to some remarks on the finitude problem and the error problem. These two problems indicate two obstacles that the dispositionalist must overcome in order to meet the extensionality condition.
在第 2 节中,我们概述了规范性条件可能如何在意义账户中施加约束的若干方式。我们建议,在所有这些方式中,这个条件至少在表面上对意向主义意义论提出了压力。在本节的其余部分,我们将限制讨论有限性问题和错误问题的一些评论。这两个问题指出了意向主义者必须克服的两个障碍,以满足外延性条件。

Blackburn responds to the finitude problem by pointing out that familiar dispositional properties (such as fragility) are in a sense infinitary: “there is an infinite number of places and times and strikings and surfaces on which it could be displayed” (1984 [2002: 35]). If a glass has infinitary dispositions, perhaps a human does too, and perhaps this will yield an extended disposition that covers the case of queries involving inaccessible numbers. Even though we have no way of getting an ordinary glass to Alpha Centauri (it would decay long before it got there), we can think of it as possessing an extended disposition to break there: breaking is what the glass would be disposed to do were its dispositions on earth allowed to manifest themselves on Alpha Centauri. Likewise, even though Jones has no disposition to answer queries involving inaccessible numbers, responding with the sum is what Jones would be disposed to do were her dispositions in accessible cases allowed to manifest themselves in inaccessible cases. This would in turn allow us to say that the answer that she would accept in those cases is “the one that would be given by reiterating procedures [Jones is] disposed to use, a number of times” (1984 [2002: 35]).
Blackburn 通过指出熟悉的倾向性属性(如脆弱性)在某种意义上是无限的来回应有限性问题:“它可以展示的地方、时间、打击和表面有无限多”(1984 [2002: 35])。如果一个玻璃具有无限的倾向性,也许人类也是如此,也许这将产生一个涵盖涉及无法访问的数字的查询案例的扩展倾向。即使我们没有办法让普通玻璃到达半人马座(它在到达那里之前就会腐烂),我们可以将其视为具有在那里破碎的扩展倾向:如果玻璃在地球上的倾向性得以在半人马座上展现,那么破碎就是玻璃被倾向于做的事情。同样,即使琼斯没有回答涉及无法访问数字的查询的倾向,但用总和回答是琼斯在可访问案例中的倾向性得以在无法访问案例中展现时会做的事情。 这反过来可以让我们说,在那些情况下,她接受的答案是“重复 [琼斯所] 倾向使用的程序,多次给出的答案”(1984 [2002: 35])。

Blackburn’s response to the finitude problem is open to criticism. First, Blackburn’s talk of procedures Jones is disposed to use is illegitimate in this context: to “use” a “procedure” is to follow a rule, and we cannot help ourselves to the idea that Jones is following the rule for addition here (or any rule, for that matter), as it is Jones’s status as a rule-follower that we are hoping to recover from facts about her dispositions. What we can say is that as far as the accessible cases go, the answers Jones is disposed to give conform to the rule for addition. But, of course, they also conform to the rule for quaddition. What makes Jones an adder, and not a quadder, according to Blackburn’s suggestion, is that were Jones’s dispositions in the accessible cases allowed to manifest themselves in the inaccessible cases, she would respond with the sum, and not the quum.
Blackburn 对有限性问题的回应是容易受到批评的。首先,Blackburn 对 Jones 可能使用的程序的谈论在这个语境中是不合法的: "使用" "程序" 是遵循一条规则,我们不能帮助自己接受 Jones 正在遵循这里的加法规则(或者任何规则),因为我们希望从关于她倾向的事实中恢复她作为规则追随者的地位。我们可以说的是,就可接近的情况而言,Jones 倾向于给出的答案符合加法规则。但是,当然,它们也符合 quaddition 规则。根据 Blackburn 的建议,使 Jones 成为加法者而不是 quadder 的是,如果 Jones 在可接近的情况下的倾向得以在不可接近的情况中表现出来,她会回应和而不是 quum。

However, Boghossian (2015: 341) points out that there is a crucial disanalogy between this case and the extended disposition to break on Alpha Centauri plausibly ascribed to the glass. To think of the dispositions the glass has on earth as manifesting themselves on Alpha Centauri, we don’t need to think of the glass in any way that is inconsistent with its nature as a physical object. It can be regarded as having the same intrinsic physical characteristics on Alpha Centauri as it has on earth, and if it is true that, given those characteristics, it would break if struck on Alpha Centauri, that suffices for the attribution of the extended disposition to break on Alpha Centauri. Matters stand differently with Jones. In order to think of Jones’s dispositions to respond in accessible cases as manifesting themselves in inaccessible cases, we would have to think of her in a way that is inconsistent with her nature as a finite biological being. This is because responding to queries involving inaccessible numbers would require, let’s suppose, a brain the size of the universe. But the fact that with a brain the size of the universe the sum would be produced no more warrants the attribution of the relevant extended disposition to Jones than does the fact that with a brain the size of the universe she would outplay Magnus Carlsen warrant the attribution to her of the potential to win the world chess championship. Jones has no extended disposition of the sort adumbrated by Blackburn.[10] The upshot, then, is that Jones’s dispositions do not determine whether she means plus rather than some quus-like function by ‘+’, where quus diverges from plus for inaccessible numbers.[11]
然而,博吉西安(2015: 341)指出,在这种情况与玻璃被赋予延伸到阿尔法半人马座的打破倾向之间存在着关键的不类比。要想象玻璃在地球上具有的性质在阿尔法半人马座上表现出来,我们无需以任何与其作为物理对象的本质相矛盾的方式来思考玻璃。可以认为它在阿尔法半人马座上具有与在地球上相同的固有物理特性,如果事实上,根据这些特性,如果在阿尔法半人马座上受到打击,它会破裂,那么这就足以归因于在阿尔法半人马座上延伸的打破倾向。而琼斯的情况则不同。为了认为琼斯对可接触案例的反应倾向在不可接触的情况下表现出来,我们必须以与其作为有限生物存在的本质相矛盾的方式来思考她。这是因为假设对涉及不可接触数字的查询做出反应需要一个宇宙大小的大脑。 但事实是,即使拥有一个像宇宙般大小的大脑,这个总和产生的事实并不能令人更多地将相关的广义倾向归因于琼斯,就像即使拥有一个像宇宙般大小的大脑,她能击败马格努斯·卡尔森的事实也不能令人归因于她有赢得世界象棋冠军的潜力。琼斯没有像布莱克本所描绘的那种广义倾向。 [10] 因此,琼斯的倾向并不决定她是否通过“+”意味着加号而不是某种类似 quus 函数,其中 quus 在不可访问的数字上与加号不同。 [11]

We’ve followed Boghossian (2015) in setting up the finitude problem as fundamentally a problem about determinacy. In a recent paper, Jared Warren admits that solving the finitude problem, thus construed, turns on solving the error problem (2020: 268), and proceeds to offer an attempted solution to that problem. Consider the following proposal: the fact which constitutes Jones’s meaning addition by ‘+’ is the fact that, when faced with arithmetical queries involving the ‘+’ sign, Jones is stably disposed to reply with the sum in the overwhelming majority of normal situations.[12] What are normal situations, and what is it for a disposition to be stable? Normal situations are those in which neither external nor internal factors are interfering with Jones’s general cognitive functioning. More specifically, the normal situations are those in which Jones is clearheaded—situations in which the air is not permeated with mind-bending chemicals, in which Jones is not drunk, exhausted, or badly hungover, so that neither external causes nor internal causes are interfering with her cognitive performance. Furthermore, to say that Jones’s disposition to respond with the sum is stable is to say that, as the number of arithmetical queries she has faced increases, the ratio of answers that give something other than the sum to answers that give the sum tends towards zero. And to keep the bar relatively low, we don’t require that in normal conditions it is metaphysically impossible for Jones to answer with something other than the sum. We only require that, when such conditions obtain, it is rational to be nearly certain that she will answer with the sum. Call the disposition which we have described here disp. disp corresponds to the meaning-constituting dispositions that Warren proposes as offering a solution to the error problem. The proposal is intended to be reductive. Warren notes that “normalcy”, defined as he defines it, “isn’t semantic or intentional or otherwise problematically question-begging” (2020: 271).
我们跟随了博戈西安(2015 年)将有限性问题设置为根本上关于确定性的问题。在最近的一篇文章中,贾里德·沃伦承认,因此构建的解决有限性问题取决于解决错误问题(2020: 268),并试图提出对该问题的解决方案。考虑以下建议:构成琼斯意思的“+”的事实是,当面对涉及“+”号的算术查询时,在绝大多数正常情况下,琼斯稳定地回答总和。什么是正常情况,什么是稳定的倾向?正常情况是指既没有外部因素也没有内部因素干扰琼斯一般认知功能的情况。更具体地说,正常情况是指琼斯头脑清醒的情况——在这些情况下,空气没有被令人迷惑的化学物质弥漫,琼斯没有喝醉、筋疲力尽或者宿醉,因此既没有外部原因也没有内部原因干扰她的认知表现。 此外,说琼斯对答案为总和的倾向是稳定的,意味着随着她面对算术查询的次数增加,给出总和以外答案的比例趋近于零。为了保持标准相对较低,我们并不要求在正常情况下琼斯不可能以总和以外的答案作答。我们只要求,当存在这样的条件时,合理地几乎可以肯定她会以总和作答。我们称这里描述的倾向为 disp。disp 对应于沃伦提出的构成意义的倾向,作为解决错误问题的方案。该提议旨在还原。沃伦指出,“正常性”,如其定义的那样,“不是语义上的,也不是有意向的,或者其他方式上有问题的前提”(2020: 271)。

However, Warren’s attempt to solve the error problem can be questioned. The error problem arises as a result of the fact that the following two possibilities are consistent with Jones’s possession of disp. First, Jones means addition by ‘+’ and is responding correctly to the relevant queries. Second, Jones means some quus-like function and is responding incorrectly. What Warren thus needs is a characterisation of normal situations such that the latter possibility is ruled out. Thus, what is required is a characterisation of normal situations such that, when those situations obtain, we are entitled to be nearly certain that Jones will answer with the sum, and that in answering with the sum Jones is responding correctly. The trouble is that there is an infinite range of functions F1,…, Fn that have different extensions from the addition function. If Jones means some function Fi among them, and if she answers with the sum, she would be answering incorrectly. Thus, the normal situations have to be such that their obtaining ensures that Jones means by ‘+’ none of the functions in this open-ended and infinite set. The question that drives Kripke’s Wittgenstein’s objection is: how could this be achieved other than through the inclusion of a clause in the characterisation of normal situations to the effect that Jones means addition by ‘+’ (or at least, that Jones means a function with the same extension as addition)? How could the obtaining of a non-semantically characterised set of situations have the effect of excluding every member of an open-ended and infinite set of semantically or intentionally characterised states of affairs (Jones means F1 by ‘+’, Jones means F2 by ‘+’, and so on ad infinitum)?[13]
然而,对于 Warren 试图解决错误问题的尝试可能会受到质疑。错误问题的出现是因为以下两种可能性与 Jones 拥有 disp 的事实是一致的。首先,Jones 通过‘+’的意思是加法,并且对相关的查询做出了正确回应。其次,Jones 的意思是某种类似 quus 的函数,并且做出了错误的回应。因此,Warren 所需要的是一种对正常情况的描述,以便排除后一种可能性。因此,所需的是一种对正常情况的描述,使得当这些情况发生时,我们几乎可以肯定 Jones 会用总和来回答,并且在回答总和时 Jones 是正确的。问题在于存在着无限范围的函数 F1 ,…, Fn ,它们从加法函数有不同的延伸。如果 Jones 的意思是其中的某个函数 Fi ,并且她用总和来回答,那么她将会回答错误。因此,正常情况必须是这样的,它们的发生可以确保 Jones 通过‘+’的意思不是这个开放且无限集合中的任何函数。 克里普克对维特根斯坦的异议的核心问题是:除了在对正常情境的表征中包含这样一个条款,即琼斯通过‘+’表示加法(或至少是琼斯表示与加法相同延伸的函数),还能以何种方式实现?如何使得非语义化特征的情境集能排除无限的语义化或故意性特征状态的所有成员(琼斯通过‘+’表示 F1 ,琼斯通过‘+’表示 F2 ,以及无穷尽的其他情况)? [13]

Thus, it can be argued that the dispositionalist account offered by Warren either fails to resolve the indeterminacy problem or does so only at the expense of deploying semantic and intentional notions, which is inconsistent with its reductive aspirations.[14]
因此,可以认为 Warren 提出的性情主义解释要么未能解决不确定性问题,要么只能通过运用语义和意图概念来解决,而这与其简化野心不一致。

Postscript to section 4: Lewis on Natural Properties
第四节的后记:关于自然属性的刘易斯

A reductionist position that has been somewhat neglected in the rule-following literature is suggested by David Lewis in his “New Work for a Theory of Universals” (1983). Lewis writes:
在关于遵循规则的文献中被忽略了一种还原主义立场,这种立场由大卫·刘易斯在他的《普遍论理论的新工作》(1983 年)中提出。刘易斯写道:

The naive solution is that adding means going on in the same way as before when the numbers get big, whereas quadding means doing something different; there is nothing present in the subject that constitutes an intention to do different things in different cases; therefore he intends addition, not quaddition. We should not scoff at this naive response. It is the correct solution to the puzzle. But we must pay to regain our naiveté. Our theory of properties must have adequate resources to somehow ratify the judgement that instances of adding are all alike in a way that instances of quadding are not. The property of adding is not perfectly natural, of course, not on a par with unit charge or sphericality. And the property of quadding is not perfectly unnatural. But quadding is worse by a disjunction. So quaddition is to that extent less of a way to go on doing the same, and therefore it is to that extent less of an eligible thing to intend to do. (1983: 376)
天真的解决方案是,在数字变大时,添加意味着继续以与以前相同的方式进行,而四次意味着做一些不同的事情;主体中没有任何存在的东西构成在不同情况下做不同事情的意图;因此,他打算进行加法,而不是四次方。我们不应嘲笑这种天真的回应。这是解开谜题的正确方法。但我们必须付出代价来恢复我们的天真。我们的属性理论必须有足够的资源,以某种方式验证这样的判断:加法的实例在某种方式上都是相同的,而四次方的实例则不是。加法属性当然不是完全自然的,当然,不如单位电荷或球形那样自然。而四次方的属性并非完全不自然。但四次方通过一个析取更差。因此,在某种程度上,四次方在继续以相同方式进行方面较少,因此在某种程度上,意图进行的事情也较少。

Take a predicate like ‘green’. The totality of facts about our previous use and dispositions to use ‘green’ are consistent with it referring to the property green but also with it referring to the property grue. So, what might ground the claim that the property green is somehow privileged as the referent of ‘green’? Lewis can be taken to advocate a form of “interpretationism” according to which semantic facts are constitutively determined by the best theory of the data (J. R. G. Williams 2007). Among the a priori constitutive constraints governing what counts as the best theory is a principle requiring that the referents assigned to expressions be the most natural of those consistent with the data. Since green is more natural than grue, it is more “eligible” than grue to be assigned to ‘green’ as its referent. Likewise for adding and quadding. In this way, the indeterminacy left open by facts about use is fended off, Lewis thinks.
采取像“绿色”这样的谓词。关于我们先前使用和使用倾向的所有事实的总体是一致的,它既可以指向绿色这个属性,也可以指向蓝绿色这个属性。那么,是什么可能支持这样一种说法,即绿色这个属性在“绿色”一词的指称中具有某种特权?根据路易斯的观点,可以认为他提倡一种“解释主义”,即语义事实在本质上是由数据的最佳理论决定的(J. R. G. 威廉姆斯,2007 年)。在统治什么构成最佳理论的先验构成约束之中,有一个原则要求分配给表达式的指称物应该是与数据一致的最自然的。由于绿色比蓝绿色更自然,它比蓝绿色更“合适”地被指定为“绿色”这个词的指称物。同样适用于添加和四重。路易斯认为,通过这种方式,由使用所留下的不确定性被抵消。

Lewis’s proposal is not ad hoc, as the notion of a natural property that it utilises is required, for example, by his account of laws of nature (see the entry on David Lewis, section 4.6). However, its application to the rule-following problem faces a number of challenges. First, it is not obvious how it extends to the mathematical examples that are the focus of Kripke’s Wittgenstein. Boghossian writes,
Lewis 的提议并非临时的,因为它利用的自然属性的概念是必需的,例如,正如他对自然法则的解释所需(参见关于 David Lewis 的条目,第 4.6 节)。然而,将其应用于遵循规则问题面临着许多挑战。首先,不明显的是它如何延伸到克里普基的维特根斯坦专注的数学例子。博戈西安写道:

I see no obvious notion of naturalness that will cover both the notion of a natural property, as it might figure in an account of similarity or lawlikeness, and that of a natural function. (2015: 355)
我看不到一个明显的自然性概念,可以涵盖自然属性的概念,它可能在相似性或类似性的解释中起作用,以及自然功能的概念。 (2015: 355)

It has also been argued that even if Lewis’s proposal might meet the extensionality condition, it cannot meet the normativity condition (Merino-Rajme 2015).
有人争论说,即使 Lewis 的提议可能符合外延性条件,但它无法满足规范性条件(Merino-Rajme,2015 年)。

Lewis’s proposal is also likely to be challenged on epistemological grounds similar to those used by Kripke and Wright in dismissing the suggestion that quaddition can be ruled out in light of the fact the hypothesis that Jones meant quaddition is less simple than the hypothesis that he meant addition. A speaker can know that in response to the query 68+57, ‘125’ is the answer that accords with what she meant by ‘+’, without having to infer this from facts about her previous linguistic behaviour. That is to say, in recognising that the answer ‘125’ fits what we mean by ‘+’, we do not proceed “by inference to the best semantic explanation of [our] previous uses of that expression” (Wright 2001: 109; see also Kripke 1982: 40). But this is apparently what we would have to do if the “simplicity” suggestion were correct: the best explanation would be yielded by the simplest of the hypotheses consistent with our previous linguistic behavior. The “simplicity” suggestion thus apparently makes a mystery of our (generally) non-inferential semantic knowledge. Lewis’s suggestion will be challenged on similar grounds. We do not infer what we mean by ‘+’ from facts about naturalness together with constitutive principles governing interpretation. Again, the account appears to make a mystery of the non-inferential nature of much of our semantic knowledge. Moreover, it faces difficulties in accommodating the authority normally credited to self-ascriptions of meaning. For Lewis, in virtue of the role they play in his account of scientific laws, simplicity and naturalness are objective notions. However, a speaker’s opinions about what she means, unlike, say, her opinions about the structure of the world or of our hypotheses about it, are generally authoritative, unless there are special reasons to doubt them. What might the basis for this default authority be, if what she means is determined by facts about simplicity and naturalness?
刘易斯的建议在认识论上也可能面临挑战,类似克里普克和赖特用来驳斥这样一种观点:即根据“简单性”建议,认为琼斯的意思是“四加法”可以被排除,因为这一假设比他的意思是“加法”更为简单。说话者可以知道,在回答“ 68+57 ”的问题时,“125”是符合她所理解的“+”的答案,而不必从她先前的语言行为中推论出来。换句话说,我们在认识到答案“125”符合我们对“+”的理解时,并没有“通过推理得出对之前使用该表达式的最佳语义解释”(赖特 2001:109;参见克里普克 1982:40)。然而,如果“简单性”建议是正确的,这显然是我们必须做的:最佳解释将由与我们先前的语言行为一致的假设中最简单的假设产生。因此,“简单性”建议似乎使我们的(一般来说)非推理语义知识成为一个谜。刘易斯的建议将在类似的理论基础上受到挑战。 我们并不通过关于自然性的事实以及解释性原则来推断我们对“+”的意义。再次,这种说法似乎使我们语义知识的非推理性质成为一个谜。此外,它在适应通常归因于意义自述的权威方面面临困难。对于刘易斯来说,由于它们在他关于科学定律的阐述中的角色,简单性和自然性是客观的概念。然而,说话者关于她意思的看法,不像她对世界结构或我们对其假设的看法那样,通常是具有权威性的,除非有特殊理由怀疑它们。如果她的意思是由简单性和自然性的事实确定的,那么这种默认权威的基础可能是什么?

Moreover, the Lewisian view seems to be a form of semantic platonism (Child 2011: 126), in so far as it upholds the idea that our meaning what we do by our words is somehow guaranteed by the structure of reality. But it might be taken to be a radical form of semantic platonism, in so far as it seems to leave no room for the speaker’s contribution to the singling out of properties (see section 3.3). Unlike other versions of semantic platonism, it is vulnerable to a complaint to the effect that the subjective perspective of the thinker or speaker is entirely annihilated.
此外,Lewisian 观点似乎是一种语义学柏拉图主义形式(Child 2011: 126),因为它坚持我们用言辞表达的意义在某种程度上是由现实的结构所保证的观念。但它可能被视为一种激进的语义学柏拉图主义,因为它似乎没有留下发言者对属性单一化的任何空间(参见第 3.3 节)。与其他版本的语义学柏拉图主义不同,它容易受到一种批评,即思想者或说话者的主观视角完全被抹杀的影响。

For a lucid exposition and critique of Lewis’s position, see J.R.G. Williams (2007). For rare examples of treatments of Lewis’s views in the context of the rule-following literature, see Merino-Rajme (2015), Glüer (2017), and Azzouni (2017).
对于对刘易斯立场的清晰阐述和批评,请参见 J.R.G.威廉姆斯(2007 年)。关于在规则遵循文献背景下处理刘易斯观点的少见例子,请参见 Merino-Rajme(2015 年)、Glüer(2017 年)和 Azzouni(2017 年)。

5. Non-Reductionism 5. 非简化论

The apparently very serious problems we outlined for the dispositionalist conception of meaning have been taken by a number of philosophers to show that we ought to resist the temptation to explain meaning and content in more basic terms. How might one formulate a non-reductionist position? On Stroud’s view, it amounts to denying that we can explain
我们为倾向性意义观所勾勒出的看似非常严重的问题,已被一些哲学家认为表明我们应该抵制诱惑,不要用更基本的术语来解释意义和内容。如何制定一种非还原主义立场?在斯特劳德的观点中,这意味着否认我们可以解释。

the phenomena of meaning and understanding “from outside” them, as it were, without attributing intentional attitudes or supposing that anything means anything or is understood in a certain way to those whose understanding is being accounted for. (Stroud 2000: viii)
理解和意义的现象“从外部”看来,不归因于故意的态度或假设任何事物意味着任何事情或以某种方式理解的人们。 (Stroud 2000: viii)

More generally, we might say that the facts constitutive of the semantic domain cannot be characterised or explained in non-semantic terms, that is, without employing the notions of meaning or understanding; the facts constitutive of rule-following cannot be characterised or explained without employing the notion of rule-following. Some philosophers who embrace non-reductionism also defend the view according to which semantic facts do not supervene on anything; they are metaphysically fundamental. Boghossian relies on the finitude problem to argue that, if meaning facts are determinate, then they cannot supervene on non-semantic facts (Boghossian 2015). However, the denial of supervenience is not essential to non-reductionism (cf. Child 2019b): at the core of the position is the idea that any attempt to account for meaning in more basic terms is hopeless or philosophically confused.
更普遍地说,我们可以说构成语义领域的事实不能用非语义术语来描述或解释,也就是说,不能不使用意义或理解的概念;构成遵循规则的事实不能在不使用遵循规则的概念的情况下进行描述或解释。一些支持非归纳主义的哲学家还辩护说,语义事实并不以任何东西为监督;它们在形上上是基础的。博戈辛依赖于有限性问题来论证,即如果意义事实是确定的,那么它们就不能依赖于非语义事实(博戈辛,2015 年)。然而,否认监督并不是非归纳主义的核心(参见 Child,2019b):该立场的核心是任何试图用更基础的术语解释意义的尝试是毫无希望或在哲学上混淆的想法。

Kripke briefly considers the possibility that the states of meaning or understanding, or the facts about meaning and understanding, are primitive or sui generis, which he cashes out as the idea that
克里普克简要考虑了意义或理解状态,或者关于意义和理解的事实可能是原始的或自成一体的可能性,他将其解释为

meaning addition by “plus” … is simply a primitive state, not to be assimilated to sensations or headaches or any “qualitative” states, nor to be assimilated to dispositions, but a state of a unique kind of its own. (1982: 51)
通过“加法”…的含义仅仅是一种原始状态,不能归类为感觉、头痛或任何“定性”状态,也不能归类为倾向,而是一种独特的状态。 (1982: 51)

He raises two complaints against this approach. First, he characterizes it as desperate, insofar as “it leaves the nature of this postulated primitive state … completely mysterious” (1982: 51), for such an approach does not provide an account of what makes it possible for one to “be confident that [one] does, at present” mean what one does (1982: 51). Second, he thinks that a non-reductionist account does not address the “logical difficulty implicit in Wittgenstein’s sceptical argument” (1982: 51), which is that it would seem that we could not “conceive of a finite state which could not be interpreted in a quus-like way” (1982: 52).
他对这种方法提出了两点批评。首先,他将其描述为拙劣的,因为“它让这个被假定的原始状态的性质完全成为了神秘”(1982 年:51),因为这种方法没有解释是什么使得一个人“确信[自己]目前所意味的”(1982 年:51)。其次,他认为非还原主义的解释没有解决“维特根斯坦怀疑论论证中隐含的逻辑困难”(1982 年:51),即似乎我们无法“构想一个不能用类 quus 方式解释的有限状态”(1982 年:52)。

Some philosophers claim that Kripke’s treatment of the non-reductionist position is unsatisfactory. McGinn, who appears to ignore Kripke’s brief discussion of non-reductionism, thinks that there is an
一些哲学家声称,克里普克对非还原主义立场的处理是不令人满意的。麦金似乎忽视了克里普克对非还原主义的简要讨论,他认为

undefended and undisclosed premise [in the sceptic’s argument], namely that semantic discourse cannot be regarded as irreducible. (1984: 82)[15]
未经辩护且未披露的前提[在怀疑者的论证中],即语义讨论不能被视为不可简化。 (1984: 82)

McGinn also notes that Kripke has no qualms with adopting a non-reductionist view of meaning in other works—see, for instance, Kripke 1972: 94–97. Goldfarb thinks that “the conception Kripke exploits is basically physicalistic” (1985 [2002: 95]), and thus that pursuing a non-physicalist approach hasn’t been ruled out. Boghossian thinks that what Kripke needs for his treatment of non-reductionism to succeed is an argument from queerness aiming to show that there is something inherently queer about meaning properties, which he fails to provide (1989 [2002: 180]; cf. Hattiangadi 2007: 47–50). Ultimately, Boghossian is sympathetic to the non-reductionist approach, though he thinks that “it really is not plausible” that such a conception might be true of linguistic meaning (1989 [2002: 179]); on his view, it is facts about mental content that are irreducible.
麦金还指出,克里普克在其他作品中对意义采取非约化观点并不感到不安—例如,参见克里普克 1972 年:94–97 页。戈德法布认为,“克里普克利用的概念基本上是物理主义的”(1985 [2002: 95]),因此追求非物理主义方法并未被排除。博戈西安认为,克里普克在其对非约化主义的处理中成功所需的是一个从奇异性出发的论证,旨在显示意义属性本质上有些奇怪,而他未能提供这样的论证(1989 [2002: 180];参见哈廷加迪 2007 年:47–50 页)。最终,博戈西安对非约化主义方法持有同情态度,尽管他认为“这种概念真的不太可能”适用于语言意义(1989 [2002: 179]);在他看来,不可简化的是关于心理内容的事实。

Before exploring several of the non-reductionist positions proposed in recent years, we should note that some of the proponents of non-reductionism think that Kripke’s sceptical challenge is based on confusion, and that our task is to unearth that confusion. Thus, on their view, the proper response is not to solve the sceptical problem by showing that the sceptic failed properly to acknowledge some set of facts (or some features of some such facts), but to dissolve it by showing that there is, in fact, no problem. McDowell, for instance, argues that Kripke misunderstands the dialectic pursued by Wittgenstein in Philosophical Investigations. On his view,
在探讨近年来提出的几种非还原主义立场之前,我们应该注意到,一些非还原主义者认为克里普克的怀疑挑战是基于混淆的,并且我们的任务是揭示这种混淆。因此,在他们看来,正确的回应不是通过展示怀疑论者未能正确承认某组事实(或某些事实的某些特征)来解决怀疑问题,而是通过展示实际上并不存在问题来消解它。例如,麦克道尔认为克里普克误解了维特根斯坦在《哲学研究》中所追求的辩证法。在他看来,

the right response to the paradox, Wittgenstein in effect tells us, is not to accept it but to correct the misunderstanding on which it depends, (1984 [1998: 229])
对于这个悖论,维特根斯坦实际上告诉我们的正确反应并不是接受它,而是纠正它依赖的误解。(1984 [1998: 229])

which puts us in a position to dissolve the paradox and, with it, the problem of how meaning is possible. This involves renouncing the problematic assumption that understanding an expression requires interpreting that expression. McDowell’s diagnosis isn’t confined to linguistic expressions. What we ought to resist is the thought that,
这使我们有能力化解这个悖论,从而解决意义如何可能的问题。这涉及放弃一个问题化的假设,即理解一个表达需要解释这个表达。麦克道尔的诊断不仅局限于语言表达。我们应该抵制的是这样的想法,

whatever is in a person’s mind at any time, it needs interpretation if it is to sort items outside the mind into those that are in accord with it and those that are not. (1992 [1998: 268])
一个人在任何时候心中所思所想,如果要将心外的事物分类为与之一致或不一致的,就需要解释。 (1992 [1998: 268])

Similarly, Stroud thinks that the paradox is “an expression of an unsatisfiable demand” (1990 [2000: 88]), namely, the demand for
同样,斯特劳德认为这个悖论是“一种无法满足的需求的表达”(1990 年[2000: 88]),即对

some facts, the recognition of which would not require that we already speak and understand a language, and some rules, which would tell us what, given those facts, it was correct to say.
一些事实,认识这些事实并不要求我们已经能够说和理解一种语言;还有一些规则,告诉我们基于这些事实,如何正确表达。

Such facts and such rules would have to be such as to “serve to get us into language in the first place” (Stroud 1990b [2000: 94]). The demand is most strongly manifested in Kripke’s assumption that there must be an item that instructs or tells the speaker what to do with her expressions. Stroud claims that proper engagement with Wittgenstein’s remarks reveals this assumption to be misguided, for it embarks us on a regress (Stroud 1996 [2000: 180–185]). Once we recognize the misguidedness of the picture that Kripke assumes, we will no longer feel the force of the sceptical challenge, Stroud thinks. However, contra Stroud, it seems that the extensionality condition is all we need in order to pose a stripped-down version of the question that the sceptic is raising, namely, the question of what makes the standards of correctness that govern our uses of expressions and our deployments of concepts possible.
这样的事实和这样的规则必须是“首先帮助我们进入语言”的(Stroud 1990b [2000: 94])。“需求最强烈地体现在克里普克的假设中,即必须有一个项目指示或告诉说话者如何处理她的表达。史特劳德声称,通过适当地参与维特根斯坦的言论,可以揭示这一假设是错误的,因为它使我们陷入了回归(Stroud 1996 [2000: 180–185])。一旦我们认识到克里普克所假设的图像是错误的,史特劳德认为,我们将不再感受到怀疑的挑战。然而,与史特劳德相反,似乎外延条件就足以提出怀疑者提出的问题的简化版本,即我们的表达和概念使用的正确性标准是如何可能的问题。

It might be argued that the urgency of this question, or even its very intelligibility, is merely a symptom of a frame of mind from within which meaning seems impossible, and that our proper task is to leave this confused frame of mind behind (e.g., McDowell 1992 [1998: 272, 274]; 1998a: 57–58). However, this line of reasoning might be taken to presume that a question is urgent or intelligible only if it is in principle possible to offer a reductive answer to it, that constructive philosophy is necessarily reductive philosophy. Verheggen argues that the rejection of reductionism does not commit one to quietism. As she sees it, non-reductionism can be constructive; it can revolve around advancing and defending positive claims (Verheggen 2000, 2003). Even though the project of providing necessary and sufficient conditions for meaning is hopeless, the non-reductionist can still aspire to articulate necessary conditions that aren’t “even remotely trivial” (Myers and Verheggen 2016: 3), and to draw illuminating connections between meaning and other irreducible phenomena. Through a creative reconstruction of Davidson’s triangulation argument, Verheggen argues that interaction with a second individual and aspects of the shared world is a necessary condition for one’s having a language and thoughts (see chapter 1 of Myers and Verheggen 2016). The constitution of the standards of correctness that govern language and thought necessitates that the individual be aware of the possibility of being mistaken, an awareness that can only be grounded in linguistic interactions with another individual and features of the world. Thus, the triangulation argument is taken by Verheggen to reveal the hopelessness of the reductive ambition, insofar as it shows that we cannot offer an account of the constitution of the standards of correctness that govern language and thought without presupposing that a commitment to those standards is already in place. (See also Sultanescu and Verheggen 2019 for an account of the Davidsonian answer to Kripke’s sceptical challenge and Verheggen 2017a for an account of the Davidsonian answer to Wittgenstein’s paradox.) William Child also defends a variety of non-reductionism, one that “does not give merely pleonastic answers but aims to say something genuinely informative” (Child 2019a: 97). He does so by relying on Wittgenstein’s remarks on meaning and rule-following.

According to Wright, “it is an important methodological precept that we do not despair of giving answers to constitutive questions too soon” (2001: 191). He goes on to propose a non-reductionist view that is sensitive to the difficulty of accounting for our knowledge of what we mean. (As we noted in section 2, Wright takes Kripke’s sceptic to impose a legitimate epistemological constraint on answers to the skeptical challenge.) He offers a “judgment-dependent” account of intention, according to which what one intends is determined by one’s best judgment about what one intends. Insofar as the concepts of intention and meaning are “relevantly similar” (2001: 206), this account can also claim to shed light on the nature of meaning. Thus, we might say, on Wright’s behalf, that what one means by an expression is determined by one’s best judgment about what one means. The notion of judgment is taken by Wright as primitive. Still, the authoritative nature of first-personal avowals is allegedly vindicated. See especially essays 5–7 in Wright 2001.[16]

Wright and Boghossian recently offered independent arguments to the effect that the adoption of a non-reductionist view of meaning does not secure the intelligibility of the idea of guidance by a rule. According to Wright, the only way in which rule-following can be understood is if it conforms to what he calls “the modus ponens model” (2007: 491). The model states that an act is a genuine instance of rule-following if it can be rationalized by, on the one hand, citing the rule and, on the other, indicating that the circumstances, which must be specifiable without appealing to the rule, call for its application. Wright investigates basic cases of language use, and argues that if we assume that the modus ponens model applies to these cases, we are saddled with an Augustinian picture, according to which conceptual capacities are necessarily prior to capacities to use language. At the same time, he accepts that Wittgenstein has revealed the bankruptcy of the Augustinian picture in Philosophical Investigations; so, the idea of antecedent conceptual capacities is unintelligible. This is taken to show that the modus ponens model cannot apply to basic cases. Wright calls this “the minor premise problem”, and argues that it compels us to accept that “in the basic case we do not really follow—are not really guided by—anything” (2007: 497).

Boghossian also relies on the modus ponens model to argue that a non-reductionist position does not allow us to make sense of the intelligibility of rule-following. He offers “an intuitive characterization” of the phenomenon of rule-following, according to which it has the following structure:

a state that can play the role of rule-acceptance; and some non-deviant causal chain leading from that state to a piece of behavior that would allow us to say that the rule explains and (in the personal-level case) rationalizes the behavior in question. (2012: 31)

What allows us to say that the rule explains and rationalizes the behaviour is an act of inference from the rule to what it requires in particular contexts. But inference is, according to Boghossian, “an example of rule-following par excellence” (2012: 40), which indicates that the act of inference must fit the intuitive characterization above, thus requiring a further act. A regress is unavoidable for the proponent of non-reductionism, according to Boghossian. This is the inference problem.

Arguably, the two problems are anticipated in Kripke’s line of reasoning, especially in his remark, mentioned earlier in this section, to the effect that the non-reductionist view faces a logical difficulty (1982: 51–52). The difficulty seems to arise from three claims that seem uncontroversial but are inconsistent:

  1. the state of meaning plus by ‘+’ must guide the speaker in her applications of the expression ‘+’;
  2. a state of meaning can be interpreted in more than one way;
  3. something that can be interpreted in more than one way cannot guide.

We might think that one’s state of meaning something by an expression is not the sort of thing that one can interpret (thus denying (ii)), or that whether something can be interpreted in more than one way is irrelevant for the question of whether it can guide (thus denying (iii)). But neither of these options entitles us to reject (i); the non-reductionist still owes us an account of what it is to be guided by a rule (or by one’s understanding of an expression).

Miller offers an account of guidance by drawing on McDowell’s writings on Wittgenstein. He argues that the inference problem and the minor premise problem are not genuine difficulties for the non-reductionist. On his view, the upshot of Wittgenstein’s reflections on rule-following is that

in applying a rule R in a particular case there need be no further inferential step—over and above that involving R itself—mediating between acceptance of R and that particular application. (2015b: 405)

This is just what it is for rule-following not to be a matter of interpretation: the rule is applied immediately, as it were—without the mediation of a further rule, such as an inference rule, in the manner suggested by Boghossian, or a rule for the deployment of a prior concept, in the manner suggested by Wright. What puts an agent in a position to follow a rule is her having been trained into a practice or custom of following rules of that sort (2015b: 407), where the notions of training, practice, and custom are semantically characterized and cannot receive further philosophical illumination.

Thus, according to some defenders of non-reductionism, rule-following has an essentially social character. One dispute that is intramural to the non-reductionist approach concerns the precise way in which this social character should be understood. When it comes to meaning, Verheggen distinguishes between communitarian views, according to which

having a (first) language essentially depends on meaning by one’s words what members of some community mean by them,

and interpersonalist views, according to which

having a (first) language essentially depends on having used (at least some of) one’s words to communicate with others, (Myers and Verheggen 2016: 84; see also Verheggen 2006 for the initial articulation of the distinction)

and goes on to defend the interpersonalist view. Relatedly, there is the question of how we should understand the notions of practice and custom and the role that they play in a correct conception of intentionality, broadly construed. It might be thought, as Wittgenstein seems to suggest in Philosophical Investigations (e.g., #198; #201–202), that it is essential to thinking and speaking that one be trained into practices or customs, and thus that our conception should reflect the centrality of these notions (McDowell 1984, M. Williams 1999, Stroud 2000, M. Williams 2010, Miller 2015b); at least initially, this appears to favour the communitarian view. (See also Section 4 of Haase 2018 for a different kind of attempt to flesh out the Wittgensteinian notion of practice, and Pettit 1990 for a form of non-reductionism on which communal interaction is required in at least some cases of rule-following). But on the Davidsonian conception, at the centre of which is disagreement and the need to settle it rationally against the constraints of the shared world, the idea of practice might not serve any explanatory purpose; even though shared beliefs may be essential for thought, shared standards of correctness are not essential for language. The Davidsonian conception is a variety of interpersonalism. This intramural debate is very much ongoing.

Some philosophers have explored the possibility of a middle path between reductive dispositionalism and non-reductionism. Thus, in a paper in which she discusses Stroud’s view, Ginsborg distinguishes between austere forms of non-reductionism, which she takes to be incompatible with constructive philosophising about meaning, and “less austere and partly reductionist” approaches, which allow that “we could account for meaning in terms of a more basic idea of goal-directed human activity” (2011a: 153), but without allowing that such activity can be captured in purely dispositionalist or physicalist terms. She goes on to articulate a less austere view, which explains meaning in terms of a notion of normativity that she takes to be primitive. On this view, for someone to mean something by an expression is for her to have a disposition to apply it in particular contexts and, crucially, to take the manifestations of that disposition to be appropriate. Taking one’s responses to be appropriate

does not depend on the antecedent grasp of a rule or standard determining that response as correct rather than incorrect, or even on the awareness that there is such a rule or standard; (2011a: 169)

this establishes the primitive status of the notion of appropriateness. Thus, Ginsborg provides a dispositionalist account of meaning, albeit with a crucial proviso to the effect that the relevant dispositions are to be characterized in normative terms. She thinks that the account can serve as a straight solution to the sceptical challenge. On the one hand, it purports to vindicate the normativity of meaning, and thus to meet the normativity condition; on the other hand, it purports to account for the distinction between correct and incorrect applications of expressions, and thus to meet the extensionality condition. For Ginsborg, the set of correct applications of an expression are the applications that one is disposed to regard as appropriate—the applications that ought, in the primitive sense, to be made (see Ginsborg 2011b for a more detailed account of her solution to Kripke’s problem).

Ginsborg’s view is in some respects similar to Robert Brandom’s. Brandom seeks to explain meaning in terms of use, where use is specified in a way that is

neither so generous as to permit semantic or intentional vocabulary, nor so parsimonious as to insist on purely naturalistic vocabulary. (1994: xiii)

Thus, his approach to meaning might also be viewed as less austere and partly reductionist. We do not have room to explain the details of Brandom’s intricate view; suffice it to say that, on that view, the proper specification of the use that determines meaning essentially involves normative vocabulary. The facts that determine meaning are facts about the entitlements and commitments that are implicit in the performances of speech acts. Thus, what a speaker means by an utterance is to be understood as consisting, roughly, in the performances that she is committed to in virtue of the utterance as well as in the performances that entitle her to make it. Ultimately, these facts are “products of human activity” (xiv), being a matter of our adopting normative attitudes toward one another--of taking one another to be committed or entitled, in light of our performances, to various other performances. However, although there are similarities, there are also important differences between Brandom’s approach and Ginsborg’s. While for Brandom the norms that are constitutive of meaning are socially instituted, for Ginsborg they are natural. Moreover, the notion of appropriateness that Ginsborg fleshes out is more basic than the notion of reason, and thus more basic than the notions of entitlement and commitment that Brandom takes to be constitutive of meaning (Ginsborg 2011a: 172fn21). Still, Ginsborg does think that “expressions have meanings only in virtue of there being ways in which they ought to be applied” (2012: 132). So, to put it crudely, on both accounts, meaning facts are reduced to facts or considerations about what ought to be the case. The question that the austere non-reductionist will raise is whether the latter kinds of fact are apt to solve the indeterminacy problem, and thus to meet the extensionality condition.

The challenge of the sceptic makes perspicuous the fact that a non-semantically described pattern is compatible with indefinitely many interpretations. Does the appeal to the normative domain help us rule out the sceptic’s alternative hypotheses? Given that, arguably, on a partly reductionist picture of the sort that Brandom proposes, utterances are, ultimately, nothing more than “normatively constrained noise- or mark-makings” (Whiting 2006: 11), that is, non-semantically described performances that stand in normative relations to other non-semantically described performances, it does not seem that we have the resources to single out as privileged a particular interpretation or standard of correctness. Similarly, a normative pattern that instantiates how one ought, in the primitive sense proposed by Ginsborg, to go on with respect to an expression seems to be consistent with more than one semantic interpretation of that expression. It might be thought that Ginsborg could appeal to the non-semantically characterized disposition in order to fix the meanings of the relevant expression. However, we have already shown in section 4 that dispositionalist accounts face very serious obstacles (see Verheggen 2015; Chapter II of Myers and Verheggen 2016, and Miller 2019 for more discussion of the failure of dispositionalism in relation to Ginsborg’s view). So, the proponent of the less austere approach to meaning owes the austere non-reductionist an account of how the extensionality condition might be met. (See Haddock 2012 and Sultanescu 2021 for more discussion of Ginsborg’s view, and Rosen 1997, McDowell 2002, Hattiangadi 2003, and Whiting 2006 for more discussion of Brandom’s view.)

Bibliography

  • Azzouni, Jody, 2017, The Rule-Following Paradox and its Implications for Metaphysics. Springer International Publishing AG.
  • Blackburn, Simon, 1984 [2002], “The Individual Strikes Back”, Synthese, 58(3): 281–301; reprinted in Miller and Wright 2002: 28–44 (ch. 3). doi:10.1007/BF00485244
  • Boghossian, Paul A., 1989 [2002], “The Rule-Following Considerations”, Mind, 98(392): 507–549; reprinted in Miller and Wright 2002:141–87 (ch. 9). doi:10.1093/mind/XCVIII.392.507
  • –––, 1990, “The Status of Content”, Philosophical Review, 99(2): 157–184; reprinted in Boghossian 2008: 51–70 (ch. 2). doi:10.2307/2185488
  • –––, 2005, “Is Meaning Normative?” Philosophy–Science–Scientific Philosophy, Paderborn: Mentis, 205–218.
  • –––, 2008, Content and Justification: Philosophical Papers, Oxford: Clarendon Press.
  • –––, 2012, “Blind Rule-Following”, in Mind, Meaning, and Knowledge: Themes From the Philosophy of Crispin Wright, Annalisa Coliva (ed.), Oxford: Oxford University Press, 27–48 (ch. 1).
  • –––, 2015, “Is (Determinate) Meaning a Naturalistic Phenomenon?”, in Meaning Without Representation: Essays on Truth, Expression, Normativity, and Naturalism, Steven Gross, Nicholas Tebben, and Michael Williams (eds.), Oxford: Oxford University Press, 331–358. doi:10.1093/acprof:oso/9780198722199.003.0016
  • –––, forthcoming, “The Normativity of Meaning Revisited”, in Meaning, Decision, and Norms: Themes from the Work of Allan Gibbard, Billy Dunaway and David Plunkett (eds.), Ann Arbor, MI: Maize Books.
  • Brandom, Robert, 1994, Making It Explicit: Reasoning, Representing, and Discursive Commitment, Cambridge, Mass.: Harvard University Press.
  • Bridges, Jason, 2014, “Rule-Following Skepticism, Properly So Called”, in Varieties of Skepticism: Essays after Kant, Wittgenstein, and Cavell, James Conant and Andrea Kern (eds.), Berlin: De Gruyter, 249–288. doi:10.1515/9783110336795.249
  • –––, 2016, “Meaning and Understanding”, in Glock and Hyman 2016: 375–389. doi:10.1002/9781118884607.ch23
  • Byrne, Alex, 1996, “On Misinterpreting Kripke’s Wittgenstein”, Philosophy and Phenomenological Research, 56(2): 339–343. doi:10.2307/2108524
  • Child, William, 2001, “Pears’s Wittgenstein: Rule-Following, Platonism, Naturalism”, in Wittgensteinian Themes: Essays in Honour of David Pears, David Charles and William Child (eds.), Oxford: Oxford University Press, 81–114 (ch. 4).
  • –––, 2011, Wittgenstein, New York: Routledge. doi:10.4324/9780203817759
  • –––, 2019a, “‘We Can Go No Further’: Meaning, Use, and the Limits of Language”, in Wittgenstein and the Limits of Language, Hanne Appelqvist (ed.), New York: Routledge, ch. 4.
  • –––, 2019b, “Meaning, Use, and Supervenience”, in Wittgenstein on Philosophy, Objectivity, and Meaning, James Conant and Sebastian Sunday (eds.), Cambridge: Cambridge University Press, 211–230. doi:10.1017/9781108151764.012
  • Churchland, Paul M., 1981, “Eliminative Materialism and the Propositional Attitudes”, Journal of Philosophy, 78(2): 67–90. doi:10.2307/2025900
  • Davidson, Donald, 1984, Inquiries into Truth and Interpretation, Oxford: Clarendon Press.
  • –––, 2001, Subjective, Intersubjective, Objective, Oxford: Clarendon Press.
  • Davies, David, 1998, “How Sceptical Is Kripke’s ‘Sceptical Solution’?”, Philosophia, 26(1–2): 119–140. doi:10.1007/BF02380061
  • Ebbs, Gary, 2016, “Rules and Rule-Following”, in Glock and Hyman 2016: 390–406. doi:10.1002/9781118884607.ch24
  • Field, Hartry H., 1980, Science Without Numbers: A Defence of Nominalism, Oxford: Blackwell and Princton, NJ: Princeton University Press.
  • –––, 1989, Realism, Mathematics and Modality, Oxford: Blackwell.
  • Fodor, Jerry A., 1990, A Theory of Content and Other Essays, Cambridge, MA: MIT Press.
  • Fogelin, Robert J., 1976 [1987], Wittgenstein, London: Routledge. Second edition, 1987.
  • Gampel, Eric H., 1997, “The Normativity of Meaning”, Philosophical Studies, 86(3): 221–242. doi:10.1023/A:1017967412131
  • Gibbard, Alan, 2012, Meaning and Normativity, Oxford: Oxford University Press. doi:10.1093/acprof:oso/9780199646074.001.0001
  • –––, 2018, “Responses to Hattiangadi, Wikforss, and Williamson”, Inquiry, 61(7): 767–782. doi:10.1080/0020174X.2018.1502072
  • Ginsborg, Hannah, 2011a, “Inside and Outside Language: Stroud’s Nonreductionism about Meaning”, in The Possibility of Philosophical Understanding: Reflections on the Thought of Barry Stroud, Jason Bridges, Niko Kolodny, and Wai-hung Wong (eds.), New York: Oxford University Press, ch 8.
  • –––, 2011b, “Primitive Normativity and Skepticism about Rules”, Journal of Philosophy, 108(5): 227–254. doi:10.5840/jphil2011108518
  • –––, 2012, “Meaning, Understanding and Normativity”, Aristotelian Society Supplementary Volume, 86: 127–146. doi:10.1111/j.1467-8349.2012.00211.x
  • –––, 2018, “Leaps in the Dark: Epistemological Skepticism in Kripke’s Wittgenstein”, in Skepticism: Historical and Contemporary Inquiries, G. Anthony Bruno and A.C. Rutherford (eds.), London: Routledge, ch. 8.
  • –––, 2020, “Wittgenstein on Going On”, Canadian Journal of Philosophy, 50(1): 1–17. doi:10.1017/can.2019.48
  • –––, forthcoming, “Going on as One Ought: Kripke and Wittgenstein on the Normativity of Meaning”, Mind & Language, first online: 8 March 2021. doi:10.1111/mila.12342
  • Glock, Hans-Johann and John Hyman (eds.), 2016, A Companion to Wittgenstein, Chichester, UK: John Wiley & Sons. doi:10.1002/9781118884607
  • Glüer, Kathrin, 1999, “Sense and Prescriptivity”, Acta Analytica, 14(23): 111–128.
  • –––, 2017, “Rule-Following and Charity: Wittgenstein and Davidson on Meaning Determination”, in Verheggen 2017b: 46–68 (ch. 4). doi:10.1017/9781316145364.005
  • Glüer, Kathrin and Peter Pagin, 1998, “Rules of Meaning and Practical Reasoning”, Synthese, 117(2): 207–227. doi:10.1023/A:1005162503125
  • Glüer, Kathrin and Åsa Wikforss, 2009, “Against Content Normativity”, Mind, 118(469): 31–70. doi:10.1093/mind/fzn154
  • –––, 2018 [2020], “The Normativity of Meaning and Content”, in The Stanford Encyclopedia of Philosophy (Fall 2020 Edition), Edward N. Zalta (ed.), URL = <https://plato.stanford.edu/archives/fall2020/entries/meaning-normativity/>.
  • Goldfarb, Warren, 1985, “Kripke on Wittgenstein on Rules”, Journal of Philosophy, 82(9): 471–488; reprinted in Miller and Wright 2002: 92–107 (ch. 6). doi:10.2307/2026277
  • –––, 2012, “Rule-Following Revisited”, in Wittgenstein and the Philosophy of Mind, Jonathan Ellis and Daniel Guevara (eds.), New York: Oxford University Press, 73–90. doi:10.1093/acprof:oso/9780199737666.003.0005
  • Grice, H. Paul, 1989, Studies in the Ways of Words, Cambridge, MA: Harvard University Press.
  • Guardo, Andrea, 2012, “Kripke’s Account of the Rule-Following Considerations: Kripke’s Account of Rule-Following Considerations”, European Journal of Philosophy, 20(3): 366–388. doi:10.1111/j.1468-0378.2010.00414.x
  • Haase, Matthias, 2009, “The Laws of Thought and the Power of Thinking”, Canadian Journal of Philosophy Supplementary Volume, 35: 249–297. doi:10.1080/00455091.2009.10717650
  • –––, 2018, “The Representation of Language”, in Language, Form(s) of Life, and Logic: Investigations after Wittgenstein, Christian Martin (ed.), Boston: De Gruyter, 219–250,.
  • Haddock, Adrian, 2012, “Meaning, Justification, and ‘Primitive Normativity’”, Aristotelian Society Supplementary Volume, 86: 147–174. doi:10.1111/j.1467-8349.2012.00212.x
  • Hale, Bob, 2017, “Rule-Following, Objectivity, and Meaning” in Hale, Wright, and Miller 2017: 619–648. doi:10.1002/9781118972090.ch24
  • Hale, Bob, Crispin Wright, and Alexander Miller (eds.), 2017, A Companion to the Philosophy of Language, second edition, Chichester, UK: John Wiley & Sons. doi:10.1002/9781118972090
  • Handfield, Toby and Alexander Bird, 2008, “Dispositions, Rules, and Finks”, Philosophical Studies, 140(2): 285–298. doi:10.1007/s11098-007-9148-2
  • Hanks, Peter, 2017, “Predication and Rule-Following”, in Philosophy and Logic of Predication, Piotr Stalmaszczyk (ed.), (Studies in Philosophy of Language and Linguistics, 7), Frankfurt am Main: Peter Lang, 199–223.
  • Hattiangadi, Anandi, 2003, “Making It Implicit: Brandom on Rule Following”, Philosophy and Phenomenological Research, 66(2): 419–431. doi:10.1111/j.1933-1592.2003.tb00269.x
  • –––, 2006, “Is Meaning Normative?”, Mind & Language, 21(2): 220–240. doi:10.1111/j.0268-1064.2006.00312.x
  • –––, 2007, Oughts and Thoughts: Rule-Following and the Normativity of Content, Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199219025.001.0001
  • –––, 2017, “The Normativity of Meaning”, in Hale, Wright, and Miller 2017: 649–669. doi:10.1002/9781118972090.ch25
  • –––, 2018, “The Normativity of meaning and the Hard Problem of Intentionality”, Inquiry, 61(7): 742–754. doi:10.1080/0020174X.2018.1424524
  • Horwich, Paul, 1998, Meaning, Oxford: Clarendon Press. doi:10.1093/019823824X.001.0001
  • –––, 2010, Truth-Meaning-Reality, Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199268900.001.0001
  • –––, 2012, Wittgenstein’s Metaphilosophy, Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199588879.001.0001
  • –––, 2015, “Kripke’s Wittgenstein”, in Meaning Without Representation: Expression, Truth, Normativity, and Naturalism, Steven Gross, Nicholas Tebben, and Michael Williams (eds.), Oxford: Oxford University Press, 359–376. doi:10.1093/acprof:oso/9780198722199.003.0017
  • Jackman, Henry, 2003, “Foundationalism, Coherentism and Rule-Following Scepticism”, International Journal of Philosophical Studies, 11(1): 25–41. doi:10.1080/0967255032000050420
  • Kremer, Michael, 2000, “Wilson on Kripke’s Wittgenstein”, Philosophy and Phenomenological Research, 60(3): 571–584. doi:10.2307/2653815
  • Kripke, Saul A., 1972, Naming and Necessity, Cambridge, MA: Harvard University Press.
  • –––, 1982, Wittgenstein on Rules and Private Language, Cambridge, MA: Harvard University Press.
  • Kusch, Martin, 2006, A Sceptical Guide to Meaning and Rules: Defending Kripke’s Wittgenstein, Montreal: McGill-Queen’s University Press.
  • Lewis, David, 1983, “New Work for a Theory of Universals”, Australasian Journal of Philosophy, 61(4): 343–377. doi:10.1080/00048408312341131
  • Mackie, J. L., 1977, Ethics: Inventing Right and Wrong, Harmondsworth: Penguin.
  • Martin, Charles Burton, 1994, “Dispositions and Conditionals”, The Philosophical Quarterly, 44(174): 1–8. doi:10.2307/2220143
  • Martin, Charles Burton and John Heil, 1998, “Rules and Powers”, Philosophical Perspectives, 12: 283–312. doi:10.1111/0029-4624.32.s12.13
  • McDowell, John, 1984 [1998], “Wittgenstein on Following a Rule”, Synthese, 58(3): 325–363; reprinted in McDowell 1998b: 221–262 (ch. 11). doi:10.1007/BF00485246
  • –––, 1992 [1998], “Meaning and Intentionality in Wittgenstein’s Later Philosophy”, Midwest Studies in Philosophy, 17: 40–52; reprinted in McDowell 1998b: 263–278 (ch. 12). doi:10.1111/j.1475-4975.1992.tb00141.x
  • –––, 1998a, “Response to Crispin Wright”, in Knowing Our Own Minds, Barry C. Smith, Crispin Wright, and Cynthia Macdonald (eds.), Oxford: Clarendon Press, 47–62.
  • –––, 1998b, Mind, Value, and Reality, Cambridge, MA: Harvard University Press.
  • –––, 2002, “How Not to Read Philosophical Investigations: Brandom’s Wittgenstein”, in R. Haller and K. Puhl (eds.), Wittgenstein and the Future of Philosophy: A Reassessment after Fifty Years, Vienna: obvhpt, 245-56; reprinted in The Engaged Intellect: Philosophical Essays, Cambridge, MA.: Harvard University Press, 2009, ch. 6.
  • McGinn, Colin, 1984, Wittgenstein on Meaning, Oxford: Blackwell.
  • Merino-Rajme, Carla, 2015, “Why Lewis’ Appeal to Natural Properties Fails to Kripke’s Rule-Following Paradox”, Philosophical Studies, 172(1): 163–175. doi:10.1007/s11098-014-0282-3
  • Miller, Alexander, 2006, “Meaning Scepticism”, in The Blackwell Guide to the Philosophy of Language, Michael Devitt and Richard Hanley (eds.), Oxford, UK: Blackwell Publishing, 91–113. doi:10.1002/9780470757031.ch5
  • –––, 2010a, “The Argument from Queerness and the Normativity of Meaning”, in Wahrheit, Bedeutung, Existenz, Martin Grajner and Adolf Rami (eds.), Boston: De Gruyter, 107–124. doi:10.1515/9783110324068.107
  • –––, 2010b, “Kripke’s Wittgenstein, Factualism, and Meaning”, in The Later Wittgenstein on Language, Daniel Whiting (ed.), Palgrave Macmillan, 167–190.
  • –––, 2011, “Rule-Following Skepticism”, in The Routledge Companion to Epistemology, Sven Bernecker and Duncan Pritchard (eds.), New York: Routledge, ch. 42.
  • –––, 2012, “Semantic Realism and the Argument from Motivational Internalism”, in Prospects for Meaning, Richard Schantz (ed.), Boston: De Gruyter, 345–362.
  • –––, 2015a, “Rule Following, Error Theory and Eliminativism”, International Journal of Philosophical Studies, 23(3): 323–336. doi:10.1080/09672559.2015.1042004
  • –––, 2015b, “Blind Rule-Following and the ‘antinomy of Pure Reason”, The Philosophical Quarterly, 65(260): 396–416. doi:10.1093/pq/pqv023
  • –––, 2018, Philosophy of Language, third edition, New York: Routledge. First edition 1998.
  • –––, 2019, “Rule-Following, Meaning, and Primitive Normativity”, Mind, 128(511): 735–760. doi:10.1093/mind/fzx033
  • –––, 2020, “What Is the Sceptical Solution?”, Journal for the History of Analytical Philosophy, 8(2). doi:10.15173/jhap.v8i2.4060
  • Miller, Alexander and Crispin Wright (eds.), 2002, Rule-Following and Meaning, Montréal: McGill-Queen’s University Press. doi:10.4324/9781315710679
  • Millikan, Ruth Garrett, 1984, Language, Thought, and Other Biological Categories: New Foundations for Realism, Cambridge, MA: MIT Press.
  • Myers, Robert H., and Claudine Verheggen, 2016, Donald Davidson’s Triangulation Argument: A Philosophical Inquiry, New York: Routledge. doi:10.4324/9781315885117
  • Pears, David Francis, 1988, The False Prison: A Study of the Development of Wittgenstein’s Philosophy, Volume Two, Oxford: Clarendon Press. doi:10.1093/019824486X.001.0001
  • Pettit, Philip, 1990 [2002], “The Reality of Rule-Following”, Mind, 99(393): 1–21; reprinted in Miller and Wright 2002: 188–208 (ch. 10). doi:10.1093/mind/XCIX.393.1
  • Rosen, Gideon, 1997, “Who Makes the Rules Around Here?”, Philosophy and Phenomenological Research, 57(1): 163–171. doi:10.2307/2953786
  • Sellars, Wilfrid, 1956 [1997], “Empiricism and the Philosophy of Mind”, in Minnesota Studies in the Philosophy of Science, Volume 1, Herbert Feigl and Michael Scriven (eds.), Minneapolis, MN: University of Minnesota Press, 253–329; reprinted in Empiricism and the Philosophy of Mind, Richard Rorty (ed.), Cambridge, MA: Harvard University Press, 1997.
  • Soames, Scott, 1997, “Skepticism about Meaning: Indeterminacy, Normativity, and the Rule-Following Paradox”, Canadian Journal of Philosophy Supplementary Volume, 23: 211–249. doi:10.1080/00455091.1997.10715967
  • Stroud, Barry, 1990 [2000], “Wittgenstein on Meaning, Understanding, and Community”, in Wittgenstein—Towards a Re-Evaluation: Proceedings of the Fourteenth International Wittgenstein Symposium, Rudolf Haller and Johannes Brandl (eds.), Wien: Springer, 27–36; reprinted in Stroud 2000: 80–94 (ch. 6).
  • –––, 1996 [2000], “Mind, Meaning, and Practice”, in The Cambridge Companion to Wittgenstein, Hans D. Sluga and David G. Stern (eds.), Cambridge: Cambridge University Press, 296–319; reprinted in Stroud 2000: 170–192 (ch. 11). doi:10.1017/CCOL0521460255.010
  • –––, 2000, Meaning, Understanding, and Practice: Philosophical Essays, New York: Oxford University Press. doi:10.1093/0199252149.001.0001
  • –––, 2011 [2018], “Meaning and Understanding”, in The Oxford Handbook to Wittgenstein, O. Kuusela and M. McGinn (eds.), Oxford: Oxford University Press, 294–310; reprinted in Stroud 2018: ch. 18.
  • –––, 2018, Seeing, Knowing, Understanding: Philosophical Essays, Oxford: Oxford University Press. doi:10.1093/oso/9780198809753.001.0001
  • Sultanescu, Olivia, 2021, “Meaning Scepticism and Primitive Normativity”, Pacific Philosophical Quarterly, 102(2): 357–376. doi:10.1111/papq.12339
  • –––, forthcoming, “Meaning, Rationality and Guidance”, Philosophical Quarterly. doi:10.1093/pq/pqac004
  • Sultanescu, Olivia and Claudine Verheggen, 2019, “Davidson’s Answer to Kripke’s Sceptic”, Journal for the History of Analytical Philosophy, 7(2): 7–28. doi:10.15173/jhap.v7i2.3487
  • Travis, Charles, 2006, Thought’s Footing: A Theme in Wittgenstein’s Philosophical Investigations, Oxford: Clarendon Press. doi:10.1093/acprof:oso/9780199291465.001.0001
  • Verheggen, Claudine, 2000, “The Meaningfulness of Meaning Questions”, Synthese, 123(2): 195–216. doi:10.1023/A:1005243504897
  • –––, 2003, “Wittgenstein’s Rule-Following Paradox and the Objectivity of Meaning”, Philosophical Investigations, 26(4): 285–310. doi:10.1111/1467-9205.00204
  • –––, 2006, “How Social Must Language Be?”, Journal for the Theory of Social Behaviour, 36(2): 203–219. doi:10.1111/j.1468-5914.2006.00303.x
  • –––, 2011, “Semantic Normativity and Naturalism”, Logique et Analyse, 54 (216): 553–567.
  • –––, 2015, “Towards a New Kind of Semantic Normativity”, International Journal of Philosophical Studies, 23(3): 410–424. doi:10.1080/09672559.2015.1042005
  • –––, 2017a, “Davidson’s Treatment of Wittgenstein’s Rule-Following Paradox”, in Verheggen 2017b: 69–96 (ch. 5). doi:10.1017/9781316145364.006
  • ––– (ed.), 2017b, Wittgenstein and Davidson on Language, Thought, and Action, Cambridge: Cambridge University Press. doi:10.1017/9781316145364
  • Warren, Jared, 2020, “Killing Kripkenstein’s Monster”, Noûs, 54(2): 257–289. doi:10.1111/nous.12242
  • Weatherson, Brian, 2009 [2021], “David Lewis”, in The Stanford Encyclopedia of Philosophy (Winter2021), Edward N. Zalta (ed.), URL = <https://plato.stanford.edu/archives/win2021/entries/david-lewis/
  • Whiting, Daniel, 2006, “Between Primitivism and Naturalism: Brandom’s Theory of Meaning”, Acta Analytica, 21(3): 3–22. doi:10.1007/s12136-006-1007-9
  • –––, 2007, “The Normativity of Meaning Defended”, Analysis, 67(2): 133–140. doi:10.1093/analys/67.2.133
  • –––, 2009, “Is Meaning Fraught with Ought?”, Pacific Philosophical Quarterly, 90(4): 535–555. doi:10.1111/j.1468-0114.2009.01354.x
  • –––, 2016, “What Is the Normativity of Meaning?”, Inquiry, 59(3): 219–238. doi:10.1080/0020174X.2013.852132
  • Wikforss, Åsa Maria, 2001, “Semantic Normativity”, Philosophical Studies, 102(2): 203–226. doi:10.1023/A:1004746319850
  • –––, 2018, “Does Semantics Need Normativity? Comments on Allan Gibbard, Meaning and Normativity”, Inquiry, 61(7): 755–766. doi:10.1080/0020174X.2018.1424528
  • Williams, J. R. G., 2007, “Eligibility and Inscrutability”, Philosophical Review, 116(3): 361–399. doi:10.1215/00318108-2007-002
  • Williams, Meredith, 1999, Wittgenstein, Mind and Meaning: Towards a Social Conception of Mind, London: Routledge.
  • –––, 2010, Blind Obedience: The Structure and Content of Wittgenstein’s Later Philosophy, London: Routledge. doi:10.4324/9780203870815
  • Wilson, George M., 1994 [2002], “Kripke on Wittgenstein and Normativity”, Midwest Studies in Philosophy, 19: 366–390; reprinted in Miller and Wright 2002: 234–259 (ch. 12). doi:10.1111/j.1475-4975.1994.tb00295.x
  • –––, 1998, “Semantic Realism and Kripke’s Wittgenstein”, Philosophy and Phenomenological Research, 58(1): 99–122. doi:10.2307/2653632
  • –––, 2003, “The Sceptical Solution”, in The Legitimacy of Truth: Proceedings of the Third Meeting of Italian and American Philosophers, Riccardo Dottori (ed.), Hamburg: Lit Verlag, 171–188.
  • Wittgenstein, Ludwig, 1953 [2009], Philosophical Investigations, G. E. M. Anscombe (trans.), Oxford: Blackwell. Revised fourth edition with P.M.S. Hacker and Joachim Schulte (eds.), Chichester: Wiley-Blackwell, 2009.
  • –––, 1956 [1978], Remarks on the Foundations of Mathematics, G. H. von Wright, Rush Rhees, and G. E. M. Anscombe (eds.), G. E. M. Anscombe (trans.), Oxford: B. Blackwell. Revised edition Cambridge, MA: MIT Press, 1978
  • Wright, Crispin, 1980, Wittgenstein on the Foundations of Mathematics, London: Duckworth.
  • –––, 1984 [2001], “Kripke’s Account of the Argument Against Private Language”, The Journal of Philosophy, 81(12): 759–778; reprinted in Wright 2001: 91–115. doi:10.2307/2026031
  • –––, 1989 [2002], “Critical Notice: Wittgenstein on Meaning, by Colin McGinn”, Mind, 98(390): 289–305; reprinted as “Critical Notice of Colin McGinn’s Wittgenstein on Meaning” in Miller and Wright 2002: 108–128 (ch. 7). doi:10.1093/mind/XCVIII.390.289
  • –––, 1992, Truth and Objectivity. Cambridge, MA: Harvard University Press.
  • –––, 1995 [2003], “Truth in Ethics”, Ratio, 8(3): 209–226; reprinted in Wright 2003: 183–203. doi:10.1111/j.1467-9329.1995.tb00084.x
  • –––, 2001, Rails to Infinity: Essays on Themes from Wittgenstein’s Philosophical Investigations, Cambridge, MA: Harvard University Press.
  • –––, 2003, Saving The Differences: Essays on Themes from “Truth and Objectivity”, Cambridge, MA: Harvard University Press.
  • –––, 2007, “Rule-Following without Reasons: Wittgenstein’s Quietism and the Constitutive Question”, Ratio, 20(4): 481–502. doi:10.1111/j.1467-9329.2007.00379.x
  • Zalabardo, José L., 1997 [2002], “Kripke’s Nonnativity Argument”, Canadian Journal of Philosophy, 27(4): 467–488; reprinted in Miller and Wright 2002: 274–293 (ch. 14). doi:10.1080/00455091.1997.10717482
  • –––, 2003, “Wittgenstein on Accord”, Pacific Philosophical Quarterly, 84(3): 311–329. doi:10.1111/1468-0114.00176

Other Internet Resources

  • Rule-Following, online bibliography by Martin Kusch, University of Vienna, Oxford Bibliographies.

Acknowledgements

We’re grateful to Claudine Verheggen for helpful comments. Thanks, too, to the SEP editors and reviewers for useful feedback and assistance.

Copyright © 2022 by
Alexander Miller <alex.miller@otago.ac.nz>
Olivia Sultanescu <olivia.sultanescu@concordia.ca>

Open access to the SEP is made possible by a world-wide funding initiative.
Please Read How You Can Help Support the Growth and Development of the Encyclopedia