07.03.2013 Views

Inquisitive pragmatics

Inquisitive pragmatics

Inquisitive pragmatics

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

1 Introduction<br />

<strong>Inquisitive</strong> <strong>pragmatics</strong> ∗<br />

(work in progress)<br />

Matthijs Westera<br />

February 21, 2013<br />

Grice (1975) provided a framework for describing how we, as speakers, build<br />

what we mean - speaker meaning - on top of what we literally say - sentence<br />

meaning. Most existing accounts in Grice’s framework rely on a classical semantics<br />

to provide the sentence meaning, but in principle any semantics might<br />

do. Not surprisingly, which phenomena the Gricean framework can tackle depends<br />

on the underlying semantics. These notes present a Gricean <strong>pragmatics</strong><br />

built upon unrestricted inquisitive semantics as formulated in (Ciardelli, 2009),<br />

with the notion of entailment from (Westera, 2012). It enables an account of<br />

exhaustivity implicatures, akin in spirit to (Westera, 2012), and it enables us to<br />

model the implicatures of (some kinds of) partial answers and counter questions<br />

and of the effect of rise-fall-rise intonation.<br />

2 <strong>Inquisitive</strong> semantics<br />

Following, e.g., (Groenendijk & Roelofsen, 2009), we conceive of propositions as<br />

embodying proposals to update the common ground with one of several pieces<br />

of information, the possibilities of the proposition. Let W be a set of possible<br />

worlds, assigning truth values to the atomic formulae of a language of choice.<br />

Definition 1 (Proposition, possibility)<br />

A possibility a is a set of worlds, a ⊆ ℘W.<br />

A proposition A is a set of possibilities, A ⊆ ℘℘W, with ∅ ∈ A.<br />

That is, a proposition is a set of possibilities including at least the empty possibility<br />

∅. This reflects the fact that, as Ciardelli (2009) remarks, the empty set<br />

does not really make a contribution to how we conceive of a proposition as a<br />

∗ Many thanks to Jeroen Groenendijk, Floris Roelofsen, Ivano Ciardelli, Elizabeth Coppock<br />

and Thomas Brochhagen for helpful comments on various versions of this work. Financial<br />

support from the Netherlands Organisation for Scientific Research (NWO) is gratefully<br />

acknowledged.<br />

1


proposal, except perhaps when it is on its own. By, unlike Ciardelli, including<br />

the empty possibility everywhere, we make sure it is of no importance when<br />

it pops up in the construction of a proposition. Effectively, it functions as the<br />

wastebasket of a proposition, and this will streamline some of the definitions to<br />

come. 1<br />

Propositions are assigned to formulae of propositional logic, given in Backus-<br />

Naur Form:<br />

Definition 2 (Syntax)<br />

For p a propositional letter, formulae ϕ: ϕ ∶∶ ∣ p ∣ (ϕ ∧ ϕ) ∣ (ϕ ∨ ϕ) ∣ (ϕ → ϕ)<br />

I will treat negation as a syntactic abbreviation according to ¬ϕ ∶= (ϕ → ).<br />

For all formulae ϕ, let [ϕ] denote the proposition of ϕ. We define the semantics<br />

recursively, following Ciardelli (2009), differing only in the empty possibility for<br />

atomic formulae:<br />

Definition 3 (Unrestricted inquisitive Semantics)<br />

For p a proposition letter, ϕ, ψ formulae:<br />

1. [p] = {{w ∈ W∣w(p) = true}, ∅}<br />

2. [] = {∅}<br />

3. [ϕ ∨ ψ] = [ϕ] ∪ [ψ]<br />

4. [ϕ ∧ ψ] = [ϕ] ⊓ [ψ] (where A ⊓ B = {a ∩ b ∶ a ∈ A, b ∈ B})<br />

5. [ϕ → ψ] = {{w∣ for all a ∈ [ϕ], if w ∈ a, then w ∈ f(a)}∣f ∶ [ϕ] → [ψ]}<br />

We can conveniently use formulae also to talk about possibilities, i.e., the sets<br />

of worlds contained in propositions:<br />

Definition 4 (Possibility)<br />

For all formulae ϕ, its possibility is given by ∣ϕ∣ = ⋃[ϕ].<br />

The possibility of a formula is simply its classical set-of-worlds denotation.<br />

2.1 Entailment<br />

Following the algebraic characterisation of Westera (2012), entailment and containment,<br />

as relations between propositions, are defined as follows: 2<br />

Definition 5 (Entailment and containment) For all propositions A, B:<br />

1. B entails A, B ⊧ A, iff ∃C, A ⊓ C = B<br />

1 I have yet to explore the algebraic repercussions of this decision, left untouched in<br />

(Westera, 2012). For one, it ensures that {∅} entails all propositions, just like contradictions<br />

in classical logic would. I think this is desirable. Also, it avoids the difficult question of<br />

what proposal the empty set would embody, because here, the empty set is not a proposition<br />

to begin with.<br />

2 Westera (2012) has a special name for the containment relation, namely ‘compliance’.<br />

However, I now doubt both the usefulness of giving it a special name, and the appropriateness<br />

of this name in particular.<br />

2


2. B contains A, A ⊆ B, iff ∃C, A ∪ C = B<br />

Note that in a classical semantics, and also in restricted (or ‘basic’) inquisitive<br />

semantics, entailment and containment, defined in this way, would coincide.<br />

Here, however, they are to some extent orthogonal, because the absorption laws<br />

do not hold:<br />

Fact 1 (No absorption laws)<br />

Not generally for all propositions A, B: A ⊓ (A ∪ B) = A or A ∪ (A ⊓ B) = A.<br />

This is easily verified for [p] ≠ [p ∨ (p ∧ q)], depicted in figure 1(b) and (c),<br />

respectively.<br />

The notion of entailment is rather sparse. For instance, it is easy to see that<br />

we can never get from [p∨q] to [p], depicted in figure 1(a) and (b), via pointwise<br />

intersection, without getting the additional possibility ∣p∧q∣, which is not in [p].<br />

Hence, a disjunction is not generally entailed by one of its disjuncts. On the<br />

other hand, [p∨q] is entailed by [p∨(p∧q)], depicted in figure 1(c). This shows<br />

pq<br />

✁pq<br />

p ✁ q ✁ p ✁ q<br />

(a) p ∨ q<br />

pq<br />

✁pq<br />

p ✁ q ✁ p ✁ q<br />

(b) p<br />

pq<br />

✁pq<br />

p ✁ q ✁ p ✁ q<br />

(c) p ∨ (p ∧ q)<br />

Figure 1: Circles represent possible worlds. Shaded areas represent possibilities.<br />

that entailment is sensitive to the internal structure of a proposition. 3 How<br />

precisely it is sensitive can be seen by unpacking the definition of entailment:<br />

Fact 2 (Entailment (unpacked)) For all propositions A, B:<br />

B ⊧ A iff ∀b ∈ B, ∃a ∈ A, ∃c s.t. b = a ∩ c and ∀a ′ ∈ A, a ′ ∩ c ∈ B.<br />

Here is an intuitive paraphrase: we look at all b ∈ B, which are the possibilities<br />

we wish to construct from the conclusion, A. For any such b, we try to construct<br />

3 Notably, entailment B ⊧ A does necessarily hold if every possibility of B is included in<br />

one of A (pointwise inclusion), nor if both this and the reverse hold, that every possibility of<br />

A contains one of B. The reader may verify this by comparing the following pair:<br />

1. [(p ∧ q ∧ r) ∨ (p ∧ q ∧ ¬r)] ⊧ [p ∧ q]<br />

2. [p ∨ (p ∧ q ∧ r) ∨ (p ∧ q ∧ ¬r)] /⊧ [p ∨ (p ∧ q)]<br />

The only difference between the two, is that in the latter case, the propositions both contain<br />

the overarching possibility ∣p∣.<br />

3


it by intersecting some a ∈ A, with some c, which can be any set of worlds. But<br />

whichever c we choose, we have to intersect every other possibility a ′ ∈ A with<br />

it as well, and this must not give any undesirable side-effects, i.e., possibilities<br />

a ′ ∩ c that are not in B.<br />

For some further insight, here is a characterisation of non-entailment, obtained<br />

simply by flipping all the quantifiers to their duals, and negating the<br />

inner conjunction, now read as an implication:<br />

Fact 3 (Non-entailment) For all propositions A, B:<br />

B /⊧ A iff ∃b ∈ B, ∀a ∈ A, ∀c, ∃a ′ ∈ A s.t. if b = a ∩ c; then a ′ ∩ c ∉ B.<br />

This can be read as follows: for B to not entail A, it is sufficient if there exists<br />

one b ∈ B that we cannot construct in any way without yielding an undesirable<br />

side-effect, i.e., a possibility that is not in B. We try all combinations of any<br />

a ∈ A with any possible c, but always there is some a ′ ∈ A that spoils it, because<br />

a ′ ∩ c is a possibility not in B.<br />

2.2 Fixing non-entailment by restricting the premiss<br />

The following technical results will be directly relevant for inquisitive <strong>pragmatics</strong>.<br />

Suppose for some A, B such that B /⊧ A, we were allowed to restrict the<br />

premiss B to some set of worlds s in order to make the entailment go through.<br />

The trivial fact that we will succeed can be stated as follows:<br />

Fact 4 (Every non-entailment can be fixed)<br />

For all propositions A, B, there is an s ⊆ W such that B ⊓ {s} ⊧ A<br />

Of course this is true, since B ∩ {∅} entails everything. More usefully, the set<br />

of all restrictions s that fix an entailment can be characterised by spelling out<br />

the definition of entailment for the case with a restricted premiss:<br />

Fact 5 (Which restrictions fix a non-entailment)<br />

For all propositions A, B and s ⊆ W, B ⊓ {s} ⊧ A iff ∀b ∈ B, ∃a ∈ A, ∃c s.t.<br />

1. b ∩ s = a ∩ c; and<br />

2. ∀a ′ ∈ A, ∃b ′ ∈ B s.t. b ′ ∩ s = a ′ ∩ c.<br />

This fact is not extremely transparent. 4 It is useful to explore how a nonentailment<br />

can be fixed for some particular subclasses of propositions. First,<br />

when the premiss contains only one possibility (besides the empty set), the<br />

constraints on the restriction are as follows:<br />

4 Despite the simple algebraic foundation of entailment, and its simple definition at the level<br />

of propositions, it seems that at the level of possibilities there is no easy way to think about<br />

entailment. In general, it is not difficult to see whether an entailment holds, by checking,<br />

at the level of propositions, whether you can construct the premiss from the conclusion by<br />

pointwise intersection; likewise it is generally quite easy to spot how the premiss may be<br />

restricted to fix an entailment. It is just that, at the level of possibilities, the non-locality of<br />

entailment makes the definitions cumbersome.<br />

4


Fact 6 (Fixing entailment if B = {b, ∅})<br />

For all propositions A, B, where B = {b, ∅}: B ⊓ {s} ⊧ A iff<br />

1. ∃a ∈ A s.t. b ∩ s ⊆ a; and<br />

2. ∀a ′ ∈ A, b ∩ s ⊆ a ′ or b ∩ s ⊆ a ′<br />

Proof: Suppose the right-hand side is false, i.e., item 1. or 2. is false. If<br />

1. is false, there is no a ∈ A s.t. b∩s ⊆ a, hence b∩s cannot be created and<br />

therefore B⊓{s} /⊧ A. If 2. is false, then ∃a ′ ∈ A s.t. b∩s /⊆ a ′ and b∩s /⊆ a ′ ,<br />

which means a ′ and b properly overlap, hence a ′ will create a side-effect<br />

regardless of how b ∩ s is manufactored, and therefore B ⊓ {s} /⊧ A.<br />

Suppose the right-hand side is true. Then according to item 1., ∃a ∈ A s.t.<br />

b∩s ⊆ a, hence b∩s can be manufactured from a with some c, for instance<br />

with c = b ∩ s. Furthermore, according to item 2., ∀a ′ ∈ A, b ∩ s ⊆ a ′ or<br />

b ∩ s ⊆ a ′ , that is, all a ′ ∈ A must either contain or be disjoint with b ∩ s.<br />

This ensures that if we manufacture b ∩ s by a ∩ b ∩ s, there will be no<br />

side-effects, since for all a ′ , either a ′ ∩ b ∩ s = b ∩ s or a ′ ∩ b ∩ s = ∅.<br />

Effectively, it is required of s that it implements either the material implication<br />

from b to some a, or to its complement.<br />

This result can be extended to propositions B of which all possibilities (except<br />

the empty set) are disjoint, because in such cases all possibilities can be<br />

created independently of one another, as if they were in separate propositions:<br />

Fact 7 (Fixing entailment if all nonempty b ∈ B are disjoint)<br />

For all propositions A, B s.t. ∀b, b ′ ∈ B, if b ≠ b ′ then b ∩ b ′ = ∅:<br />

B ⊓ {s} ⊧ A iff ∀b ∈ B:<br />

1. ∃a ∈ A s.t. b ∩ s ⊆ a; and<br />

2. ∀a ′ ∈ A, b ∩ s ⊆ a ′ or b ∩ s ⊆ a ′<br />

This is just fact 6, with the two requirements imposed on all b ∈ B. Fact 6 and 7<br />

are all we need for the examples I will treat in the section below on <strong>pragmatics</strong>. 5<br />

3 <strong>Inquisitive</strong> <strong>pragmatics</strong><br />

3.1 The conversational maxims<br />

Grice (1975) argued that dialogue participants adhere to (and assume eachother<br />

to adhere to) the cooperative principle:<br />

Definition 6 (The cooperative principle (Grice, 1975)) Make your contribution<br />

such as it is required, at the stage at which it occurs, by the accepted<br />

purpose or direction of the talk exchange in which you are engaged.<br />

5 Clearly, though, they will not suffice for future applications. In a future version I hope to<br />

give similar accounts of the required restrictions when B ⊆ A or B ⊑ A (pointwise subset, i.e.,<br />

∀b ∈ B, ∃a ∈ A s.t. b ⊆ a).<br />

5


From the cooperative principle a number of maxims can be derived that more<br />

directly guide a speaker’s behaviour and a hearer’s interpretation of that behaviour.<br />

These maxims fall in a number of categories, depending on whether<br />

they pertain to the quantity of what is said (e.g., too much/too little information),<br />

the relatedness of it to previous discourse (for Grice its ‘relevance’, but I<br />

will avoid that term), the quality of what is said (e.g., its truth/falsehood), or<br />

the manner in which it is said (e.g., whether it is clear and concise). Some of<br />

an open-ended set of maxims are given below.<br />

Definition 7 (Maxims) Propose a proposition only if:<br />

1. you know it to be true; (Quality)<br />

2. you consider its possibilities to be possible; (Quality)<br />

3. relative to your knowledge state it entails the QUD; (Relation)<br />

4. it supports as many of the QUD’s possibilities as possible; (Quantity)<br />

5. its possibilities are not smaller than necessary for the above. (Quantity/Manner)<br />

In these notes I will be concerned only with the first four maxims. Each of<br />

these can be translated into a set of more formal constraints on the speaker’s<br />

epistemic state:<br />

Fact 8 (Epistemic constraints imposed by the maxims)<br />

In response to a QUD A, the following holds for a cooperative speaker, with<br />

knowledge state s, who utters B:<br />

1. s ⊆ ⋃ B (Quality)<br />

2. For all b ∈ B, a ∩ s ≠ ∅ (Quality)<br />

3. B ⊓ {s} ⊧ A (Relation)<br />

4. For all a ∈ A, if ⋃ B /⊆ a, then s /⊆ a (Quantity)<br />

The constraints imposed by the qualitative maxims are quite straightforward.<br />

For the third maxim, relation, facts 5, 6 and 7 can be used to spell out the<br />

constraint in more detail. The fourth maxim embodies the requirement that<br />

if the speaker can support a possibility, she must do so, and hence that if she<br />

doesn’t, she must not be able to. This is typically how so-called ignorance<br />

implicatures have always been derived in the literature since Grice, though with<br />

variations on where the ‘what should have been said’ comes from.<br />

Under the assumption that the speaker is cooperative, the epistemic constraints<br />

in (8) translate directly into implicatures. These implicatures are due<br />

to abiding the maxims, not, e.g., flouting them, which would yield a wholly<br />

different class of implicatures that is outside our present scope.<br />

6


3.2 Exhaustivity implicatures<br />

Among the implicatures we can account for is the exhaustivity implicature of<br />

a reply to a disjunction. For brevity, I will formulate the implicatures below<br />

in standard modal logic, with the familiar ◻ and ◇ indicating the epistemic<br />

necessity and possibility for the speaker. That is, for a speaker with knowledge<br />

state s, ◻ϕ iff s ⊆ ∣ϕ∣ and ◇ϕ iff s∩∣ϕ∣ ≠ ∅. The precise model-theoretic semantics<br />

and axioms assumed for this modality are not important for our purposes (yet).<br />

(1) a. Will John come, or Mary? (p ∨ q ∨ ¬(p ∨ q))<br />

b. John will come.<br />

Implicatures:<br />

1. ◻(John will come) (Quality)<br />

2. ◻(If John will come, then Mary will come) or<br />

◻(If John will come, then Mary won’t come) or<br />

◻(If John will come, then neither will come) (Relation)<br />

3. ◇(Mary will not come) (Quantity)<br />

4. ◻(Mary will not come) (from 1, 2 and 3)<br />

Here the quantity implicature in 3 in effect disambiguates the relation implicature<br />

in 2, together with 1. entailing that Mary won’t come, i.e., exhaustivity.<br />

Notice that the third disjunct of the relation implicature in 2 is inconsistent<br />

with the quality implicature in 1. In the next examples, I will therefore omit<br />

such disjuncts from from the start.<br />

In the same vein, we can account for the exhaustivity implicature in response<br />

to a ternary disjunction:<br />

(2) a. Will John come, or Mary, or Bob?<br />

b. John will come.<br />

Implicatures:<br />

1. ◻(John will come) (Quality)<br />

2. ◻(If John will come, then Mary will come) or<br />

◻(If John will come, then Mary won’t come), and<br />

◻(If John will come, then Bob will come) or<br />

◻(If John will come, then Bob won’t come) (Relation)<br />

3. ◇(Mary will not come), ◇(Bob will not come) (Quantity)<br />

4. ◻(Mary will not come), ◻(Bob will not come) (from 1, 2 & 3)<br />

Finally, we can derive the absense of an exhaustivity implicature in the<br />

following example:<br />

(3) a. Will John come, or Mary?<br />

b. John will come, and maybe Mary too (= ‘or John and Mary’).<br />

Implicatures:<br />

1. ◻(John will come) (Quality)<br />

2. ◇(Mary will come) (Quality)<br />

7


3. ◻(If John will come, Mary will come) or<br />

◻(If John will come, Mary won’t come) (Relation)<br />

4. ◇(Mary will not come) (Quantity)<br />

5. ◻(Mary will not come) (from 1, 3 & 4)<br />

In this case, the maxim of relation does not impose any constraints on the<br />

epistemic state, because the response already entails the QUD as such. Traditionally,<br />

the response would have been regarded as ‘canceling’ the exhaustivity<br />

implicature, but I find this terminology misleading; it suggests that the implicature<br />

was first computed, and then later left out. This is not what happens.<br />

The ‘and maybe Mary too’ makes an actual contribution to the literal meaning,<br />

as a consequence of which it entails the QUD and no exhaustivity implicature<br />

is made. 6<br />

4 Expected relatedness<br />

4.1 A more flexible maxim of relation<br />

An utterance can be perfectly fine also if the speaker is not sure whether her<br />

proposal is related (in the sense of entailing the QUD relative to her knowledge<br />

state), but merely expects it to be. Two ways in which relatedness can be<br />

expected are the following:<br />

The speaker has defeasible knowledge that she expects to hold. I.e., she<br />

expects the case under discussion to be not an exception to her rules of<br />

thumb.<br />

The speaker does not have the knowledge herself, but she expects one of<br />

the other discourse participants to provide it. That is, she thinks it is<br />

quite likely distributed knowledge.<br />

In both cases the speaker can be thought of as expecting to know something<br />

(although for sure this is not the only way, and probably not the best way, of<br />

thinking about defeasible knowledge). I will insert the possibility to rely on<br />

expected knowledge explicitly in the maxim of relation:<br />

Definition 8 (Maxim of relation (v2)) Propose a proposition only if it entails<br />

the QUD relative to your knowledge state, or relative to a knowledge state<br />

you expect to enter in the future. The latter must be marked.<br />

How it is marked may be language-dependent. In English, it is marked by risefall-rise<br />

intonation, and optionally by preceding the response with ‘well’ (but<br />

this has other uses, too):<br />

6 The reasoning scheme that leads to an exhaustivity implicature is purely semantic, and<br />

therefore it generalizes to responses to expressions with existential force, and also to some<br />

cases of embedded implicature (Westera, 2012). But instead of discussing such examples here<br />

on a case-by-case basis, I am working on some general facts that will facilitate predictions<br />

regarding natural language, and that should help comparing this <strong>pragmatics</strong> to the various<br />

exhaustivity operators that have been proposed.<br />

8


(4) Will John come to the party?<br />

Well, Mary will go... [rise-fall-rise]<br />

We will investigate several such examples below. In each case, the relation<br />

implicature is computed in the same manner as before, but because it is marked<br />

it will be further embedded under a modal operator, ⊡, that represents the<br />

speaker’s expectations. I do not wish to define a formal semantics for this modal<br />

operator here; an intuitive understanding suffices for the present purposes.<br />

4.2 Expected relatedness implicatures<br />

The following examples show the effect of the updated maxim of relation:<br />

(5) a. Will John come?<br />

b. Well, it will rain. [rise-fall-rise]<br />

Implicatures:<br />

1. ◻(It will rain) (Quality)<br />

2. ⊡(◻(if it rains John will come) or ◻(if it rains John won’t come)<br />

or ◻(if it rains contradiction))<br />

3. ◇(John will come), ◇(John won’t come) (Quantity)<br />

4. ⊡(◻(John will come) or ◻(John won’t come)) (1, 2)<br />

The relation implicature in 2 is computed according to fact 6. It is disjunctive,<br />

because there are different ways to make the utterance entail the QUD. The<br />

third disjunct cannot be the case, for the speaker knows that it is raining.<br />

Of the first two disjuncts, presumably, since people dislike rain, the speaker<br />

expects the latter, but in principle the speaker’s utterance leaves the audience<br />

guessing at what knowledge the speaker expects (indeed, the dialogue can be<br />

very naturally continued with a hearer asking for clarification: ‘okay, so what<br />

do you expect?’). Note that it can be concluded from 1 and 2 together that<br />

the speaker expects to know that John will come, or that he won’t come. The<br />

quantity implicature in 3 of course still holds: neither of the disjuncts in 4 is<br />

certain.<br />

Had the response been unmarked, i.e. pronounced with neutral falling intonation,<br />

the dotted box would have been left out, and the conclusion in 4 would<br />

have directly contradicted the quantity implicatures in 3. Hence, with falling<br />

intonation, the speaker would have violated either the maxim of relation, or the<br />

maxim of quantity. Indeed, these two options are intuitively available to make<br />

sense of such a response: either the speaker violates relation because she wants<br />

to introduce a new, unrelated topic of conversation (i.e., the weather), or the<br />

speaker violates quantity with the purpose of testing the hearer (in which case<br />

the speaker could have appended ‘so what do you conclude?’ to her utterance).<br />

By exactly the same reasoning, rise-fall-rise intonation avoids an exhaustivity<br />

implicature:<br />

(6) a. Will John come, or Mary?<br />

9


. Well, John will come. [rise-fall-rise]<br />

Implicatures:<br />

1. ◻(John will come) (Quality)<br />

2. ⊡(◻(if John will come, Mary will come) or<br />

◻(If John will come, Mary won’t come) or<br />

◻(If John will come, contradiction)) (Relation)<br />

3. ◇(Mary will not come) (Quantity)<br />

4. ⊡(◻(Mary will come) or ◻(Mary won’t come)) (1, 2)<br />

In this case, no exhaustivity is implicated for the following reason. Unlike in<br />

example (1), the same dialogue but with normal intonation, here the quantity<br />

implicature in 3 is perfectly compatible with both disjuncts expected in 4,<br />

notably also with the expected knowledge that Mary will come, because the<br />

knowledge is only expected. Therefore, unlike in (1), this possibility cannot be<br />

excluded, and no exhaustivity is implicated.<br />

It seems that counterquestions, too, may rely on defeasible or otherwise<br />

expected knowledge. Obtaining a rise-fall-rise intonation on questions may be<br />

hard, if not impossible, perhaps because it interferes with the default rising<br />

intonation for questions (but I would not be surprised if some kind of marking<br />

is in fact detectible):<br />

(7) a. Will John come?<br />

b. Will it rain?<br />

Implicatures:<br />

1. ◇(it will rain), ◇(it will not rain) (Quality)<br />

2. ⊡(◻(if rain then John) or ◻(if rain then not John), and<br />

◻(if not rain then John) or ◻(if not rain then not John)) (Rel.)<br />

3. ◇(John will come), ◇(John will not come) (Quantity)<br />

4. ⊡(◻(John won’t come) or ◻(John will come iff it rains) or<br />

◻(John will come) or ◻(John comes iff it doesn’t rain)) (1, 2)<br />

This time, the relation implicature is computed according to fact 7. I have omitted<br />

the relation implicatures that have contradictions in them, because each of<br />

those is incompatible with the quality implicature anyway. It follows in 4 that<br />

either the speaker expects there to be a relation between John’s attendance and<br />

the weather, or the speaker expects John to come or not come regardless of the<br />

weather. The latter is a bit counter-intuitive: had the speaker expected this,<br />

why would she have started about the weather? As before, what exactly the<br />

speaker’s defeasible or otherwise expected knowledge is, is left underspecified by<br />

the utterance, and hence inferring it always relies heavily on context. Presumably,<br />

we can rule out that the speaker expects the weather has no bearing on<br />

John’s attendance because, had someone known the answer to the QUD already,<br />

this person would have come forward before the weather was brought up.<br />

Finalizing, recall that each of the above examples can either be read as<br />

involving the speaker’s defeasible knowledge, or as involving the speaker’s expectation<br />

as to what other dialogue participants will be able to tell her. It is<br />

10


easy to think of the latter as involving some kind of request for the other participants<br />

to come forward with this information. At this point, I will leave this<br />

as a mere intuition; but it seems to me worthwile to have a systematic account<br />

of how expectations regarding other participants’ knowledge play a role in the<br />

formation of issues.<br />

4.3 More examples: conditional partial answers<br />

We obtain interesting results if we apply the same reasoning as above, to conditional<br />

partial answers, such as the following:<br />

(8) a. Will John come or not?<br />

b. He’ll come if it’s sunny.<br />

The speaker meaning (at the least) of such responses varies with different intonation<br />

patterns. For instance, it can mean ‘given that it is sunny, he will<br />

come’, and it can suggest a high or low likelihood attached to John’s coming.<br />

Mapping out the different intonations, contexts and speaker meanings is a work<br />

in progress. For now, I will focus on rise-fall-rise intonation, in a context where<br />

it is not known whether it is sunny, and I will ignore any suggestion of high or<br />

low likelihood.<br />

One would expect a conditional answer to raise the issue of whether its<br />

antecedent is true, but as we shall see, only a strengthened biconditional answer<br />

really achieves this. Let us look at the first example.<br />

(9) a. Will John come or not?<br />

b. He’ll come if it’s sunny. [rise-fall-rise]<br />

Implicatures:<br />

1. ◻(s → j) (Quality)<br />

2. ⊡(◻((s → j) → j) (≡ ◻(¬s → j)) and<br />

◻((s → j) → ¬j) (≡ ◻(¬j))) (Relation)<br />

3. ◇(j), ◇(¬j) (Quantity)<br />

4. ⊡(◻(j) or ◻(¬j ∧ ¬s)) (from 1, 2)<br />

Surprisingly perhaps, it is implicated in 2, according to fact 6, that the speaker<br />

expects not ‘it’s sunny’, but rather ‘he’ll come also if it’s not sunny’. Together<br />

with her own knowledge, it entails in 4 that she expects the knowledge that<br />

John will come regardless of the weather, or the knowledge that John won’t<br />

come and (or ‘because’) it’s not sunny. I think both disjuncts would certainly<br />

be fine as responses, and perhaps the predictions of inquisitive <strong>pragmatics</strong> are<br />

not bad in this respect. But still, one would like the response in (9) about the<br />

weather to somehow lead to a discussion about what the weather is like.<br />

To achieve this, we may assume that the conditional response is strengthened<br />

to a biconditional. Perhaps this is due to interpreting an implication as a causal<br />

statement and applying, as we often do, closed-world reasoning. But regardless<br />

of why exactly it happens, this strengthening is common in conversation, and<br />

we can explore its effect in inquisitive <strong>pragmatics</strong>:<br />

11


(10) a. Will John come or not?<br />

b. He’ll come iff it’s sunny.<br />

Implicatures:<br />

1. ◻(j ↔ s) (Quality)<br />

2. ⊡(◻((j ↔ s) → j) (≡ ◻(¬s → j)) and<br />

◻((j ↔ s) → ¬j) (≡ ◻(s → ¬j))) (Relation)<br />

3. ◇(¬j), ◇(j) (Quantity)<br />

4. ⊡(◻s or ◻¬s)<br />

Here, combining the relation implicature with her own knowledge, the speaker<br />

implicates in 4 that she expects the knowledge that it is sunny or the knowledge<br />

that it is not sunny. Notice that this example is almost the mirror image of<br />

example (7). There, the response was a polar disjunction regarding the weather<br />

(‘will it rain?’), and the implicature was effectively a polar disjunction of biimplications;<br />

here the response itself is a bi-implication, and the implicature is<br />

a polar disjunction regarding the weather.<br />

Finally, for completeness sake, we can explore an ‘only’-conditional in inquisitive<br />

<strong>pragmatics</strong>:<br />

(11) a. Will John come or not?<br />

b. He’ll only come if it’s sunny.<br />

Implicatures:<br />

1. ◻(j → s) (Quality)<br />

2. ⊡(◻((j → s) → j) (≡ ◻(j)) or<br />

◻((j → s) → ¬j) (≡ ◻(j → ¬s)) (Relation)<br />

3. ◇(¬j), ◇(j) (Quantity)<br />

4. ⊡(◻(j ∧ s) or ◻¬j) (from 1, 2)<br />

As with the normal conditional, the predictions of inquisitive <strong>pragmatics</strong> are a<br />

bit unexpected, but not downright wrong. It is implicated not that the speaker<br />

expects knowledge about what the weather is like, but rather that she expects<br />

the knowledge that if John comes it won’t be sunny, which, together with the<br />

literal sentence meaning, is a rather cryptic way of saying that the weather<br />

doesn’t matter, that John won’t come anyway. I suspect that plugging in a more<br />

natural form of implication may make the predictions of inquisitive <strong>pragmatics</strong><br />

regarding conditional responses more intelligible, and empirically more accurate.<br />

But for the moment, at least the strengthened biconditional works fine already.<br />

Finally, note that with a more neutral intonation (though there seem to<br />

be many variations), for each of the responses the relation implicature will be<br />

incompatible with the quantity implicature. Hence, the responses will have<br />

to be interpreted either as suggesting a new topic for a conversation (flouting<br />

relation), or as intended to test the hearer’s deductive capacities (bypassing<br />

quantity).<br />

12


5 Concluding remarks<br />

This is a work in progress. All comments are very much appreciated. My main<br />

aims for the next version will be to prove more general results on the epistemic<br />

constraints imposed by the maxim of relation, in order to make the implicatures<br />

easier to compute for examples that go beyond the ones discussed.<br />

References<br />

Ciardelli, I. (2009). <strong>Inquisitive</strong> semantics and intermediate logics. (Master<br />

Thesis, University of Amsterdam)<br />

Grice, H. (1975). Logic and conversation. In P. Cole & J. Morgan (Eds.), Syntax<br />

and semantics (Vol. 3, pp. 41–58).<br />

Groenendijk, J., & Roelofsen, F. (2009). <strong>Inquisitive</strong> semantics and <strong>pragmatics</strong>.<br />

In J. M. Larrazabal & L. Zubeldia (Eds.), Meaning, content, and argument:<br />

Proceedings of the ILCLI international workshop on semantics,<br />

<strong>pragmatics</strong>, and rhetoric.<br />

Westera, M. (2012). Meanings as proposals: a new semantic foundation for<br />

gricean <strong>pragmatics</strong>. (Presented at SemDial 2012)<br />

13

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!