An Introduction to English Semantics and Pragmatics 9781399504621

A classic introduction to the study of meaning, revitalised for a new generation of linguists Ideal for undergraduate s

821 181 872KB

English Pages 186 [192] Year 2016

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

An Introduction to English Semantics and Pragmatics
 9781399504621

Table of contents :
Contents
List of figures and tables
Preface to the third edition
1 Studying meaning
2 Sense relations
3 Nouns
4 Adjectives
5 Verbs
6 Tense and aspect
7 Modality, scope and quantification
8 Pragmatic inference
9 Figurative language
10 Utterances in context
11 Doing things with words
Suggested answers to the exercises
References
Index

Citation preview

An Introduction to English Semantics and Pragmatics

Edinburgh Textbooks on the English Language General Editor Heinz Giegerich, Professor Emeritus of English Linguistics, University of Edinburgh Editorial Board Heinz Giegerich, University of Edinburgh – General Editor Laurie Bauer (University of Wellington) Olga Fischer (University of Amsterdam) Willem Hollmann (Lancaster University) Marianne Hundt (University of Zurich) Rochelle Lieber (University of New Hampshire) Bettelou Los (University of Edinburgh) Robert McColl Millar (University of Aberdeen) Donka Minkova (UCLA) Edgar Schneider (University of Regensburg) Graeme Trousdale (University of Edinburgh) titles in the series include:

An Introduction to English Syntax, Second Edition Jim Miller An Introduction to English Phonology April McMahon An Introduction to English Morphology: Words and Their Structure Andrew Carstairs-McCarthy An Introduction to International Varieties of English Laurie Bauer An Introduction to Middle English Jeremy Smith and Simon Horobin An Introduction to Old English, Revised Edition Richard Hogg, with Rhona Alcorn An Introduction to Early Modern English Terttu Nevalainen An Introduction to English Semantics and Pragmatics Patrick Griffiths An Introduction to English Sociolinguistics Graeme Trousdale An Introduction to Late Modern English Ingrid Tieken-Boon van Ostade An Introduction to Regional Englishes: Dialect Variation in England Joan Beal An Introduction to English Semantics and Pragmatics, Second Edition Patrick Griffiths and Chris Cummins An Introduction to English Phonetics, Second Edition Richard Ogden An Introduction to English Morphology, Second Edition Andrew Carstairs-McCarthy An Introduction to English Phonology, Second Edition April McMahon An Introduction to English Semantics and Pragmatics, Third Edition Patrick Griffiths and Chris Cummins Visit the Edinburgh Textbooks on the English Language website at www​ .edinburghuniversitypress.com/series/ETOTEL

An Introduction to English Semantics and Pragmatics Third Edition Patrick Griffiths Revised by Chris Cummins

Edinburgh University Press is one of the leading university presses in the UK. We publish academic books and journals in our selected subject areas across the humanities and social sciences, combining cutting-edge scholarship with high editorial and production values to produce academic works of lasting importance. For more information visit our website: edinburghuniversitypress.com © in the first edition of this work, Literary Estate of Patrick Griffiths, 2006 © in the revisions and additional third edition material, Chris Cummins, 2023 Cover design & illustration: riverdesignbooks.com Edinburgh University Press Ltd The Tun – Holyrood Road 12(2f) Jackson’s Entry Edinburgh EH8 8PJ Typeset in 10.5/12 Janson by Cheshire Typesetting Ltd, Cuddington, Cheshire and printed and bound in Great Britain A CIP record for this book is available from the British Library ISBN 978 1 3995 0460 7 (hardback) ISBN 978 1 3995 0461 4 (paperback) ISBN 978 1 3995 0462 1 (webready PDF) ISBN 978 1 3995 0463 8 (epub) The right of Patrick Griffiths and Chris Cummins to be identified as the authors of this work has been asserted in accordance with the Copyright, Designs and Patents Act 1988, and the Copyright and Related Rights Regulations 2003 (SI No. 2498).

Contents

List of figures and tablesviii Preface to the third editionix   1 Studying meaning Overview 1.1 Sentences and utterances 1.2 Types of meaning 1.3 Semantics vs pragmatics Summary Exercises

1 1 4 5 10 14 14

  2 Sense relations 16 Overview 16 2.1 Propositions and entailment 16 2.2 Compositionality 19 2.3 Synonymy 21 2.4 Complementarity, antonymy, converseness and incompatibility23 2.5 Hyponymy 26 Summary 30 Exercises 30 Recommendations for reading 31  3 Nouns 32 Overview 32 3.1 The has-relation32 3.2 Count nouns and mass nouns 40 Summary 42 Exercises 42 Recommendations for reading 43

vi

an introduction to english semantics and pr agmatics

 4 Adjectives Overview 4.1 Gradability 4.2 Combining adjective meanings with noun meanings Summary Exercises Recommendations for reading

44 44 44 47 52 53 53

 5 Verbs Overview 5.1 Verb types and arguments 5.2 Causative verbs 5.3 Thematic relations Summary Exercises Recommendations for reading

55 55 55 57 61 64 64 65

  6 Tense and aspect Overview 6.1 Talking about events in time 6.2 Tense 6.3 Aspect Summary Exercises Recommendations for reading

66 66 66 69 73 79 79 80

  7 Modality, scope and quantification Overview 7.1 Modality 7.2 Semantic scope 7.3 Quantification Summary Exercises Recommendations for reading

82 82 82 87 90 98 98 100

  8 Pragmatic inference Overview 8.1 Some ways of conveying additional meanings 8.2 The Gricean maxims 8.3 Relevance Theory 8.4 Presuppositions Summary Exercises Recommendations for reading

101 101 103 106 112 115 119 119 120



contents vii

  9 Figurative language Overview 9.1 Literal and figurative usage 9.2 Metaphor 9.3 Metonymy 9.4 Simile 9.5 Irony 9.6 Hyperbole Summary Exercises Recommendations for reading

121 121 121 123 125 127 128 129 131 131 131

10 Utterances in context Overview 10.1 Tailoring utterances to the audience 10.2 Definiteness 10.3 Given and new material 10.4 The Question Under Discussion Summary Exercises Recommendations for reading

133 133 133 134 136 143 145 146 146

11 Doing things with words Overview 11.1 Speech acts 11.2 Indicators of speech acts Summary Exercises Recommendations for reading

148 148 148 151 159 159 160

Suggested answers to the exercises161 References175 Index179

List of figures and tables

Figures

  2.1 An example of levels of hyponymy   2.2 Hyponym senses get successively more detailed as we go down the tree   2.3 Part of the hyponym hierarchy of English nouns   4.1 The simplest case of an adjective modifying a noun is like the intersection of sets   6.1 Time relations among the events described in (6.1). Time runs from left to right: open-ended boxes indicate events with no endpoint suggested    7.1 Venn diagrams for the meanings of (7.20a) and (7.20c)

28 28 29 48 69 92

Tables

  3.1 Distinguishing between count and mass nouns 41   5.1 Examples of causative sentences with an entailment from each59   5.2 Three kinds of one-clause causative with an entailment from each 61   6.1 Two-part labels for tense–aspect combinations, with examples69   6.2 The compatibility of some deictic adverbials with past, present and future time 73   6.3 A range of sentences which all have habitual as a possible interpretation74 10.1 A selection of indefinite and definite forms in English 137

viii

Preface to the third edition

When Laura Williamson from Edinburgh University Press asked me whether I would like to produce a new edition of this textbook, my first reaction was to be flattered. But, of course, what Laura said was rich in possible interpretations. I still remember being confused by just this kind of utterance at primary school, and failing to understand that the teacher’s “Would you like to sit over there?” was actually an instruction rather than just a slightly odd question. With more experience of language under my belt, naturally I understood Laura’s question as an indirect request. But, inevitably, I also wondered whether there was a shade of meaning along the lines of ‘we need a new edition, because there is so much wrong with the existing edition’. I still don’t know whether that meaning was intended – I didn’t dare ask. But the opportunity to go back and revise a text like this is always welcome, and of course I found all sorts of places where what struck me as crystal clear six years earlier strikes me as a bit murky now. So I’ve welcomed the chance to try to filter out some of these issues. I’ve also attempted to plug a few gaps in coverage, while trying not to go on at unwarranted length. This book retains much of the structure of Patrick Griffiths’s original edition, and draws on many of the examples he introduced. Therefore, much of the credit goes to him, and to the people to whom he generously gave credit, notably Heinz Giegerich, Anthony Warner, Kenji Ueda, and Janet and Jane Griffiths. For my part, I’d also like to thank Ronnie Cann for his detailed comments on the second edition and Hannah Rohde for her comments on and support with the third edition. And, of course, thanks go to the team at Edinburgh University Press for their encouragement, professionalism and, above all, patience.

ix

1  Studying meaning

Overview

This book is about how the English language allows people to convey meanings. As the title suggests, semantics and pragmatics are the two main branches of the linguistic study of meaning. Semantics is the study of the relationship between units of language and their meaning. Pragmatics is concerned with how we use language in communication, and so it involves the interaction of semantic knowledge with our knowledge of the world, including the contexts in which we say things. Explanations of terms Numbers in bold print in the index point to the pages where technical terms, such as semantics and pragmatics in the paragraph above, are explained. To illustrate some of the differences between semantics and pragmatics, let’s consider example (1.1). (1.1) That’s what I’m talking about! Language enables us to communicate about the world. This works because there are conventional links between expressions in language and aspects of the world (real or imagined) that language users talk and write about: things, activities, and so on. We sometimes use the word denote to describe these links: I denotes the speaker, talking denotes an activity, and so on. An expression is any meaningful language unit or sequence of meaningful units, such as a sentence, a phrase, a word, or a meaningful part of a word (such as the s of cats, which denotes plurality, as opposed to the s of is, which doesn’t have its own separate meaning). 1

2

an introduction to english semantics and pr agmatics

In (1.1), that denotes something that is assumed to be obvious to the hearer in the current context of utterance: it might be an object that is physically present, or it might be a linguistic expression that has just been uttered. The expression what I’m talking about also denotes something, specifically the topic that the speaker has been discussing. And the expression ’s, in (1.1), denotes a relationship of identity between those two denotations. So one possible interpretation of (1.1) is that something which is obvious to the hearer at the moment (that) corresponds (’s) to the topic that the speaker was previously addressing (what I’m talking about). But actually there’s another possible interpretation for (1.1), because it also functions as a fixed idiomatic expression meaning something like ‘That’s excellent’ or ‘That’s impressive’ or ‘That’s great news’. In this case, that would still denote something that is obvious to the hearer in the moment, what I’m talking about would denote ‘excellent’, and ’s would denote a relationship between these two things which attaches the quality ‘excellent’ to the denotation of that (often called a predication). The important point here is that there is an interplay between semantics and pragmatics. Semantics provides a set of possible meanings, and pragmatics is concerned with the choice among these possibilities. Language users can take account of context, and use general knowledge about the way the world works, to build interpretations on this semantic foundation. In understanding (1.1), we use the context of the utterance not only to understand what the speaker means by that, but also to figure out which of the possible senses of the whole utterance is relevant: the one in which they mean that that matches a previous topic of discussion, or the one in which they just mean that that is excellent. This view fits with a way of thinking about communication that was pioneered by the philosopher H. P. Grice in the mid-twentieth century (see Grice 1989 for a collection of relevant work) and which has been highly influential within linguistics. According to this view, human communication is not just a matter of encoding a signal, sending it, and decoding it at the other end. Rather, it requires active collaboration on the part of the addressee or hearer (the person the message is directed to). The hearer’s task is to try to guess what the speaker intends to convey, and the message has been communicated when (and only when) the speaker’s intention has been recognised. (By convention, I’ll refer to the ‘speaker’ and the ‘hearer’ whether we’re actually talking about speech or written communication.) The speaker’s task is to judge what needs to be said (or written) in order for this process to work: that is, so that it’s possible for the hearer to recognise the speaker’s intention.



studying me aning 3

This way of looking at communication has several important consequences: • The same string of words can convey different messages, because – depending on the context – there will be different clues to help the hearer recognise the speaker’s intention. • Similarly, different strings of words can be used to convey the same message, because there may be lots of different ways for the speaker to give clues to their intention. • Sometimes a great deal can be communicated with very few words, because of the active participation of the hearer in recovering the speaker’s meaning. • Mistakes are possible: the hearer might arrive at the wrong interpretation. If a speaker suspects that this has happened, they may try to repair the communication by saying more. This is easy in faceto-face interactions (where the speaker can monitor the hearer’s reaction for clues that they might have been misunderstood), harder in live telephone conversations or text chats (because of the lack of feedback), harder still in emails or letters, and essentially impossible in books (where the writer may never find out whether the reader understood them as intended). The rest of this chapter introduces some other concepts that are important to the study of linguistic meaning, and indicates which later chapters develop them further. It will define a few more technical terms, but (hopefully) just enough to give you a reasonable initial grasp of semantics and pragmatics and set you up for reading introductory books in this area. In everyday life, competent users of a language generally don’t give a lot of thought to the details of what they’re doing and how it works. Linguistics researchers do: they generally operate on the assumption that there are interesting things to be learned from those details. This can appear a little obsessive at times, like in the discussion of (1.1) – why should we care what that, on its own, means? And does it make a difference that ’s does a slightly different thing in the two possible interpretations? But the general goal of linguistic analysis is often to try to bring all the unconscious knowledge that we have about language to conscious awareness. It turns out that close inspection of bits of language and instances of usage – even quite ordinary ones – can sometimes give us new insights into how language really works and, by extension, how we think.

4

an introduction to english semantics and pr agmatics

1.1  Sentences and utterances

Just as we distinguish semantics and pragmatics, we distinguish sentences and utterances. Utterances are the things that have meaning in our immediate experience as language users. (1.2) presents three examples. (1.2) a. I agree with you. b. The departments. c. All that glisters is not gold. Utterances are the raw data for our linguistic analysis. Just as we use the term “speaker” to mean the sender of a particular message, whether that’s in speech or writing, we use the term “utterance” to refer to that message (or a part of it), regardless of whether or not it is actually spoken out loud. Each utterance is unique, having been produced by a particular speaker in a particular situation, and can never precisely be repeated. Even when someone “says the same thing twice”, by repeating themselves, the two utterances differ in potentially important ways because the context has changed (it may be relevant that the previous utterance was made, for instance). It’s easy to see how the content of what a speaker means by saying (1.2a) depends crucially on the context: after all, the speaker is more likely to mean that they agree with the hearer about the particular issue under discussion at the moment than they are to mean that they always or generally agree with the hearer. Notation When it matters, I use “” double quotes for utterances, ‘’ single quotes for meanings, and italics for sentences and words that are being considered in the abstract. I also use single quotes when quoting other authors, and double quotes as “scare quotes” around expressions that are not strictly accurate but potentially useful approximations. The abstract linguistic object on which an utterance is based is a s­ entence. For instance, if (1.2b) is uttered in response to the question “Who is responsible for lecture scheduling?”, we could think of (1.2b) as being based on the sentence The departments are responsible for lecture scheduling. If it had been uttered in response to some other question, it would likely correspond to a different sentence, and bear a different meaning. In the case of the utterance of the proverb (1.2c), we could argue that the underlying sentence is straightforwardly All that glisters is not gold. We



studying me aning 5

don’t need to rely on context to figure this out. However, that sentence is itself ambiguous: it has two possible meanings, ‘Not everything that glisters is gold’ or ‘Nothing that glisters is gold’. Sometimes a speaker will signal which interpretation is correct – in this case, for instance, emphasising not might hint at the ‘not everything’ ­interpretation – but often things are not so clear-cut. (We’ll see more examples like this in Section 7.3.5.) However, even if the speaker is not giving us a clear indication of which meaning to choose, context might still help us understand which one is more likely to be appropriate. In the case of this proverb, the sense that is usually relevant is the weaker ‘not everything’ meaning – the speaker is more likely to want to convey the message ‘things may not be what they appear to be’ than to convey the message ‘things are never what they appear to be’. 1.2  Types of meaning

When we think about meanings, there are several slightly different ideas that we might want to distinguish, among them speaker meaning, utterance meaning, and sentence meaning. We can think of speaker meaning as the meaning that the speaker intends to convey when they produce an utterance. Assuming that, as hearers, we’re interested in figuring out what people are trying to tell us – which is probably for our own good, in general – we are constantly having to make informed guesses about speaker meaning. Speaker meaning, construed this way, is a private thing: only the speaker really knows what they meant to convey. Indeed, a speaker may not wish to admit that they intended to convey certain meanings. Furthermore, a speaker might not be successful in conveying a ­particular meaning – or, to put it another way, the hearer might not be successful in recovering the meaning that the speaker intended. So we might want to distinguish speaker meaning from utterance meaning, which we could think of as the meaning that an utterance could be understood as conveying when interpreted by people who know the language, are aware of the context, and have whatever background knowledge could reasonably be assumed by the speaker. Utterance meaning is, in this sense, a necessary fiction or idealisation that linguists doing semantics and pragmatics have to work with. It relies on the intuitions that we have as language users about what would be likely to happen, communicatively, as a result of a particular utterance being made under a particular set of circumstances. Because utterances are instances of sentences in use, an important first step towards understanding utterance meaning is understanding

6

an introduction to english semantics and pr agmatics

sentence meaning. I’ll take this to be the meaning that people familiar with the language can agree on for sentences considered in isolation. This is a good place to start because language users have readily accessible intuitions about sentences. I appealed to a shared intuition about sentence meaning when I said that (1.2c) was ambiguous. That was a relatively subtle example, though: lots of sentences are much more obviously ambiguous, such as (1.3a), which could mean (1.3b) or (1.3c), or (1.4a), which could mean (1.4b) or (1.4c). (1.3) a. We arranged to meet yesterday. b. Yesterday, we arranged that we would meet (on some day I am not specifying here). c. We arranged a meeting that was to take place yesterday. (1.4) a. Jane saw the guy with binoculars. b. Jane, using binoculars, saw the guy. c. Jane saw the guy who was carrying binoculars. Language users’ access to the meanings of individual words – what we call lexical meaning – is less direct. We can think of the meaning of a word as the contribution it makes to the meanings of sentences in which it appears. Of course, people know the meanings of words in their language in the sense that they know how to use the words, but this knowledge is not immediately available in the form of reliable intuitions. Speakers of English might be willing to agree that strong means the same as powerful, or that finish means the same as stop; but these judgements would be at least partly wrong, as shown when we compare (1.5a) with (1.5b), and (1.5c) with (1.5d). (1.5) a. This cardboard box is strong. b. ?This cardboard box is powerful. c. Mavis stopped writing the assignment yesterday, but she hasn’t finished writing it yet. d. *Mavis finished writing the assignment yesterday, but she hasn’t stopped writing it yet.

Notation Asterisks, like the one at the beginning of (1.5d), are used in semantics to indicate that an example is seriously problematic as far as meaning goes – in this case, the sentence is a contradiction. Question marks, like the one at the beginning of (1.5b), are used to signal less serious but still noticeable oddness of meaning.



studying me aning 7

From the above examples, we might conclude that finishing is a special kind of stopping, specifically ‘stopping after the goal has been reached’, and that strong can mean either ‘durable’ or ‘powerful’ (among other possibilities), only one of which is applicable to cardboard boxes. 1.2.1  Denotation, sense, reference and deixis

Earlier in the chapter I said that expressions in a language – sentences, words, and so on – denote aspects of the world. The denotation of an expression is whatever it denotes. For many words, their denotation is a large class of things: the noun thing itself would be an extreme example. The links between language and the world are what makes it so ­indispensable – we can use language to talk about things in the world. That being the case, it is tempting to think that the meaning of an expression is just its denotation. If I wanted to explain what window meant, I might be able to do so by uttering the word and pointing to a window. And in early childhood, our first words are probably learnt through just such a process of live demonstration and pointing, known as ostension. However, this is not plausible as a general explanation of how meaning is acquired, for several reasons including the following: • After early childhood, we usually acquire word meanings through the use of language rather than the use of ostension (“A sash window is a window which you can open by sliding it upwards”). • Ostension isn’t a very good way of specifying what we’re talking about. If I just said window while pointing at the window, how would the hearer know whether I meant to label the window itself, or something I could see through it, or indeed some other property that it had (‘made of glass’, ‘transparent’, etc.)? And how would the hearer know whether I meant to refer to this particular window, or windows in general? So, in practice, when we do use ostension, it is often accompanied by explanatory utterances. • There are all kinds of abstract, non-existent and relational denotations that cannot conveniently be explained by ostension: consider memory, absence, yeti, or instead of. If we can’t think of meaning as just a matter of denotation, how can we think about it? One approach is that taken in formal semantics (“formal” because it uses systems of formal logic to set out descriptions of meaning, and theories of how the meanings of expressions relate to the meanings of their parts; see Lappin 2001). In formal semantics, it is important to consider what kind of denotation we are dealing with. Count nouns, such as tree, may be said to denote sets of things; mass

8

an introduction to english semantics and pr agmatics

nouns, like honey, denote substances; singular names denote individuals; property words, like purple, denote sets of things (the things that have the property in question); spatial relation words, like in, denote pairs of things that are linked by that spatial relation; simple sentences such as Amsterdam is in Holland denote facts or falsehoods; and so on. What is of interest is the fact that the denotation is what it is – an individual, or a set of things, or a set of pairs of things, and so on. An alternative approach takes sense to be the central concept: this is a slightly elusive idea, but we can think of it as the aspects of the meaning of an expression that give it the denotation it has. Differences in sense therefore lead to differences in denotation. Ambiguous words – of which there are many in English – have multiple possible senses, and consequently might denote different things. Consider (1.6), as part of a job advertisement. (1.6) a. The applicant should list his three most recent employers. b. The applicant should list her three most recent employers. c. The applicant should list their three most recent employers. One sense of his in (1.6a) could be loosely expressed as ‘of the most contextually salient man’, while an alternative sense could be expressed as ‘of the most contextually salient person’. Given that, in this context, his is clearly supposed to mean ‘of the applicant’, the former sense results in the inference that application is restricted to men, whereas the latter sense does not. Assuming that the advertiser does not wish to suggest that there might be such a restriction, they would probably be better advised to use an alternative formulation, such as (1.6c). The idea of sense as the core of linguistic meaning offers a helpful route into the study of semantics. This book will focus on this approach, presenting it in a version that will hopefully also form a reasonable foundation for anyone who later chooses to learn formal semantics. There are different ways in which we might try to write down “recipes” for the denotations of words. One way of doing this is in terms of sense relations, which are semantic relationships between the senses of expressions. The idea here is that, if we can explain the interconnections between words using well-defined sense relations, then it is possible for a person who knows the denotations of some words to develop an understanding of the meanings (senses) in the rest of the system. To take a trivial example, if I know that I like coriander, and I learn that cilantro means the same as coriander, I know that I like cilantro. This approach meshes well with the observation that we commonly use language to explain meanings. Chapter 2 will start to



studying me aning 9

explain the sense relations that can hold between words, and phrases, in a more systematic way. We can usefully distinguish the idea of sense from that of reference. Reference is what speakers do when they use expressions – which we call referring expressions – to pick out particular entities – which we call referents – for their audience. These referents can be of many kinds, including, for instance (with sample referring expressions in parentheses), people (“my students”), things (“the Parthenon Marbles”), times (“midday”), places (“the city centre”) and events (“her party”). Reference is strongly reliant on context. Consider (1.7), where the speaker intends to use Edinburg to refer to the city of that name in Indiana. (1.7) We drove to Edinburg today. The speaker of (1.7) would have to be sure that the hearer knows that they are in Indiana, if the utterance is not to be misunderstood as referring to the Edinburg in Illinois, or the one in Texas, or even Edinburgh in Scotland. But of course there is more context-dependence in (1.7) than just this. We in (1.7) refers to the speaker plus (usually) at least one other person. Similarly, today refers to the day on which the utterance was made. Thus, in order to understand what (1.7) means, we need to know who the speaker is and when they are speaking – which will be trivial face-to-face, but not possible if we just encounter the utterance out of its original context, such as within the pages of this book. And even in the face-to-face encounter it may not be obvious precisely who the speaker means to refer to by we. A similar problem arises with the notice (1.8), once posted on a course bulletin board. (1.8) The first tutorial will be held next week. The notice was posted in week 1 of the academic year, but not dated, and the lecturer forgot to take it down. Some students read it in week 2 and missed the first tutorial because they quite reasonably interpreted next week to mean ‘week 3’. We refer to expressions like this, which have to be interpreted in relation to their context of utterance, as deictic expressions (or instances of deixis). Deixis is pervasive in language, presumably because we can often save effort by appealing to context in this way: it’s often easier to say she than to specify a person’s name, easier to say here than to give a clear description of the place of utterance, easier to say tomorrow than to remember the date, and so on. There are different kinds of deixis, relating to:

10

an introduction to english semantics and pr agmatics

• time:  now, soon, recently, ago, tomorrow, next week . . . • place:  here, there, two kilometres away, that side, this way, come, bring, upstairs . . . • persons and entities: she, her, hers, he, him, his, they, it, this, that . . . • discourse itself: this sentence, the next paragraph, that’s what she said, this is true . . . Our semantic knowledge of the meanings of deictic expressions guides us as to how we should interpret them in context. As always, where context is concerned, these interpretations will be guesses rather than certainties: perhaps a speaker who points to an object and says this means that specific object, but perhaps they mean some property of it, or perhaps they are referring to their hand or the pointing gesture itself. Much of language is in some sense deictic: tense, which will be discussed in Chapter 6, can also be thought of this way. Reference in general is an important topic that will recur in many of the following chapters, particularly Chapter 10. 1.3  Semantics vs pragmatics

As we’ve seen, the essential difference between sentences and utterances is that sentences are abstract, and not tied to contexts, whereas utterances are identified by their contexts. This is also an important way of distinguishing between semantics and pragmatics. If you are dealing with meaning and there is no context to consider, then you are doing semantics, but if there is a context to consider, then you are probably engaged in pragmatics. (As we shall see later in this section, there is a bit of a grey area as to whether context can sometimes intrude into semantics.) In the terms we’ve discussed so far, pragmatics is the study of utterance meaning, while semantics is the study of sentence meaning and word meaning. As an example of how semantics and pragmatics relate to one another, consider how we might interpret (1.9). (1.9) The next bus goes to Cramond. Considering this as a sentence, we can think about its sentence meaning, drawing on the semantic information that we have from our knowledge of the language. Anyone who knows English well can explain various features of the meaning of (1.9): it means that a bus goes to Cramond, and that one bus that does this is, in some sense, the ‘next bus’ (perhaps the adjacent one physically, but more usually the next one to arrive). Also, more subtly, goes to Cramond can be understood either to be making



studying me aning 11

a prediction about what the next bus will do, or stating a generalisation about where that bus habitually goes. These meanings are available without considering who might be saying or writing the words, or when or where they are doing so: essentially, no context is involved. Hence, their study falls within the domain of semantics. By contrast, if we consider (1.9) as an utterance that takes place in a particular context, we can derive a richer interpretation that goes beyond the sentence meaning (which we will sometimes refer to as the literal meaning). Suppose that a passenger steps onto a bus and asks the driver whether this bus goes to Cramond, and the driver replies by uttering (1.9). Assuming that the driver is being cooperative, we can interpret (1.9) as conveying the answer ‘no’. This understanding relies on us trying to figure out why, given the contextual and background information, the driver produced the utterance (1.9). The literal meaning of (1.9) is obviously relevant to this process, but on this occasion it doesn’t relate closely to the meaning that we end up deriving: whether or not the next bus goes to Cramond doesn’t really tell us anything about whether this bus does so, and the utterance of (1.9) doesn’t generally convey the meaning ‘no’. So the interpretation of (1.9) as ‘no’, in this case, is a contextually driven additional meaning that goes beyond what was literally stated – a type of meaning that we call an implicature. And because it relies on context, the study of implicature falls within the domain of pragmatics. There are also forms of meaning which do not fall very clearly within the scope of semantics or that of pragmatics. In (1.9), the question of which bus actually is the next bus is dependent on the time and place at which (1.9) is uttered. We have already seen that the resolution of deictic expressions depends on context in this way. In order to understand what the speaker of (1.9) is committing to, we therefore need to figure out a complete interpretation of the utterance, by using contextual information and world knowledge to work out what is being referred to (and how to understand potentially ambiguous expressions, like next in this case). The basic interpretation of a sentence with these details spelled out is sometimes termed an explicature. We could think of explicature as part of pragmatics because it is context-dependent, or part of semantics because it is essential to the meaning of the utterance in a way that implicature is not. I will try to sidestep that theoretical debate in this book, although the notion of explicature will be used in the discussion of figurative language in Chapter 9.

12

an introduction to english semantics and pr agmatics

1.3.1  A first outline of semantics

Semantics, the study of word meaning and sentence meaning abstracted away from contexts of use, is primarily a descriptive subject. It is an attempt to describe and understand the nature of the knowledge about meaning that people have as a consequence of knowing a language. It is not, however, prescriptive: that is to say, it is not about defining what words “ought to mean” or pressuring speakers into abandoning some meanings and adopting others. Nor is it about etymology, because you can know a language perfectly well without knowing its history. It may be fascinating to explore how different meanings have become associated with historically related forms – arms, armour, army, armada, ­armadillo, etc. – but that knowledge is not essential for understanding or using present-day English, so it is not covered in this book. Nor do we focus on semantic and pragmatic change, for the same reason – it is interesting that the meanings of words change over time, and that, for instance, silly originally meant ‘blessed’ and subsequently ‘innocent’, but you do not need to know this in order to understand what it means now. Having said all that, studying semantics and pragmatics is helpful in understanding some of the processes involved in historical meaning change: for instance, meanings that were once context-dependent can end up becoming part of the semantics of a word. (A case in point: armada in Spanish just meant ‘an armed force’, and its use in English to mean specifically ‘a fleet of warships’ arose because Spanish Armada denoted an armed force from Spain that happened to comprise a fleet of warships. We might want to appeal to the notion of compositionality, introduced in Section 2.2, and ideas about adjective meanings, discussed in Chapter 4, to explain how this happened.) The process of giving a semantic description of language knowledge is also different from the encyclopedia-writer’s task of cataloguing general knowledge. The words tangerine and clementine illustrate a distinction that may not be part of our knowledge of English: although an expert will be able to tell these apart, most users of English will not. But our focus in this book will be on the more abstract kinds of semantic and pragmatic knowledge that underpin our ability to use language. 1.3.2  A first outline of pragmatics

Pragmatics is concerned with characterising how we go beyond what was literally said, both in terms of what additional meanings get conveyed by a speaker and in terms of how the speaker encodes them and the hearer recovers them. A crucial basis for making pragmatic infer-



studying me aning 13

ences is the contrast between what might have been uttered and what actually was uttered. Consider (1.10), from the information provided to guests at a B&B. (1.10) Food and Drink. Breakfast is served from 7:30am to 9:00am. There is a fridge in the hallway with soft drinks and snacks. Payment for these is on an honesty basis. As no further information is provided under this heading, we are invited to infer that the establishment does not serve dinner and does not provide alcoholic drinks. However, we cannot be entirely certain about this: perhaps the proprietors simply didn’t want to promise that dinner or alcoholic drinks would be available. These pragmatic inferences do not have the same status as the information that is explicitly asserted: a guest who read (1.10) would have grounds to complain if the fridge in the hallway contained no soft drinks or no snacks, but they would not (usually) have grounds to complain if the fridge also contained beer. Another widely available pragmatic inference, often called a scalar implicature, arises when words can be ordered on a semantic scale, as for example the value-judgements excellent > good > OK. (1.11) A: What was the accommodation like at the camp? B: It was OK. Speaker A can draw an inference from B’s response, because if the accommodation had been better than merely OK, B could have used the word good or indeed excellent to describe it. As B does not do so, A can infer that the accommodation was no better than satisfactory. But, as in (1.10), B is not as committed to the accommodation being ‘no better than OK’ than they are to it being ‘at least OK’. Indeed, if B sounds surprised, the inference ‘no better than OK’ may be less readily available: B then might be assumed to mean something like ‘contrary to expectations, it was acceptable, and maybe even better than that’. We’ll revisit this kind of inference in Section 8.2.2. Pragmatic inferences of this kind occur all the time in communication: even though they are really just informed guesses, they are crucial to the smooth functioning of much of our communication. Because they are informed guesses, it is one of their defining features that they can be “cancelled”, in the sense that the speaker can typically deny a pragmatic inference without seeming to contradict themselves. In (1.11), B could continue “In fact, it was pretty good” without being self-contradictory, because what is ‘pretty good’ in this case is also ‘OK’. Pragmatics is the focus of the later chapters of this book, but will also figure in sections of most of the other chapters.

14

an introduction to english semantics and pr agmatics

Summary

Hearers (including readers) have the task of guessing what speakers (including writers) intend to communicate when they produce utterances. If the guess is correct, the speaker has succeeded in conveying the meaning. Pragmatics is about how we interpret utterances – and produce interpretable utterances – taking account of context and background knowledge. Semantics is the study of the context-independent knowledge that users of a language have about the meanings of expressions, such as words and sentences. Crucially, expressions of language relate to the world outside of language. In this book, we will explore this idea through the notion of sense and of the meaning relations that hold within a language, in ways that later chapters will make clearer. Exercises

1. Here are two sets of words: {arrive, be at/in, leave} and {learn, know, forget}. There is a similarity between these two sets, in how the words relate to one another. Can you see it? Here is a start: someone who is not at a place gets to be there by arriving; what if the person then leaves? Once you have found the similarities between the two sets, answer this follow-up question: was this a semantic or a pragmatic task? 2. A student says to the tutor “How did I do in the exam?” and the tutor replies “You didn’t fail.” What the tutor opted to say allows the student to guess at the sort of grade achieved. Do you think the grade was high or low? How confident are you about this? Briefly justify your answer. In doing this, were you doing semantics or pragmatics? 3. Pick the right lock is an ambiguous sentence. State at least two meanings it can have. How many different propositions could be involved? 4. A common myth about the word kangaroo is that it comes from an expression in an Australian language (specifically Guugu Yimithirr) meaning ‘I don’t understand you’. This was supposedly because explorers in the eighteenth century pointed to a kangaroo and asked what it was, and a local replied kangaroo. What does this story tell us about the limits of ostension? And how would we disprove the claim that this is what kangaroo originally meant? 5. An old joke concerns someone reading a sign saying Dogs must be carried on this escalator and having to wait ages for a dog to appear so that they could use the escalator. If that sign really caused people any problems, how could you add a deictic term to it and thus resolve the ambiguity?



studying me aning 15

6. Example (1.6a) is potentially problematic as an utterance because, on one interpretation of his, it suggests that only men are welcome to apply for the job. Does (1.6b) run into a similar problem? What about (1.6c)?

2  Sense relations

Overview

As mentioned in Chapter 1, we generally learn the meanings of our first few words through close encounters with the world, carefully mediated by our caregivers. But once we have a start in language, we learn the meanings of most other words through language itself. This might be through having meanings deliberately explained to us – for instance, we might be told that “tiny means ‘very small’”. It might also be because we draw inferences about meanings based on our knowledge of language. For instance, if we see the title of Gerald Durrell’s book My Family and Other Animals, we might infer from this that people can potentially be classified as a type of animal. In both of these cases, we are relying on the existence of relationships between the senses of expressions (words and phrases) within the language that we speak. In the first case, we rely on the fact that the meaning of tiny can be expressed in terms of the meanings of very and small. In the second case, we use our understanding of the meaning of other to spot the existence of a relationship between the meanings of family and animals. One task for semantics is to describe these relationships systematically, with the ultimate ambition of identifying exactly what it is that a competent speaker of a language – in this case, English – has to know about the meanings of sentences. Putting this together with the speaker’s knowledge about pragmatics – how we use sentences – we would then have a picture of what a speaker knows about linguistic meaning altogether. In this chapter, we make a start on this task, beginning in the following subsection with some technical preliminaries. 2.1  Propositions and entailment

We need to account for sentence meaning in order to develop explanations of utterance meaning, because utterances are sentences put to use. 16



sense rel ations 17

The number of sentences in a human language is potentially infinite, so we cannot analyse sentence meaning just by listing every possible sentence and its interpretation (and nor do we learn sentence meanings this way, with the exception of sentences that are fixed idiomatic expressions). Both as linguistic scholars and as language learners, we have to generalise in order to discover the principles of sentence meaning. One important generalisation is that different sentences can carry the same meaning, as in (2.1a–c). (2.1) a. Sharks hunt seals. b. Seals are hunted by sharks. c. Seals are prey to sharks. Proposition is a term for the core meaning of a sentence: we could say that (2.1a–c) express the same proposition. Propositions are therefore not tied to particular words or sentences. They have the property that they are either true or false. That is not to say that the speaker or hearer (or anyone else) has to know whether the proposition actually happens to be true or false, as far as the real world is concerned – just that it must, in principle, be either true or false. The sentences in (2.1) are declaratives: that is to say, they follow the sentence pattern on which statements (utterances that explicitly convey factual information) are typically based. It is easy to see that they express propositions, because it is possible to affirm or deny or query the truth of these statements (“Yes, that’s true”, “No, that’s a lie”, “Is that really true?”). By contrast, such responses are not appropriate to utterances such as (2.2a–b), which are based on other sentence patterns. (2.2) a. What’s your name? b. Please help me. Although (2.2a–b) don’t actually express propositions, we could still try to analyse their meaning with reference to propositions. We could say that (2.2a) relates to a proposition of the form ‘the hearer’s name is ___’, and invites the hearer to supply their name to fill in the blank. Similarly, we could say that (2.2b) presents the proposition ‘the hearer will help the speaker’ and indicates that the speaker wants that proposition to be true. But it might be more useful to think about utterances like this in terms of the social actions that speakers are trying to perform when they produce them, as we shall see in Chapter 11 when discussing speech acts. An important relation that holds between propositions is entailment, which we can write as ⇒. We say that one proposition entails another (p ⇒ q, “p entails q”) if the first proposition being true guarantees that the second is also true. Entailment is also called “logical consequence”

18

an introduction to english semantics and pr agmatics

and is one of the most fundamental concepts in logic: arguments are logically valid if and only if the starting points (premises) entail the conclusions. Although entailment holds between propositions, we often talk about it holding between sentences – which is fine as long as we make sure that we are considering each sentence with one specific propositional meaning. The examples in (2.3) illustrate some of these points. (2.3)

a. b. c. d.

Moira has arrived in Edinburgh. Moira is in Edinburgh. Moira has arrived in Edinburgh ⇒ Moira is in Edinburgh *Moira has arrived in Edinburgh and she is not in Edinburgh.

When (2.3a) is true we can be sure that (2.3b) is also true – assuming, that is, that we are talking about the same person called Moira and the same place called Edinburgh in both sentences, as otherwise the propositional meanings won’t be related. If the truth of (2.3a) guarantees the truth of (2.3b), that means that the entailment shown in (2.3c) holds. Another way of stating this is to point out that (2.3d) is a contradiction: given that (2.3a) entails (2.3b), we cannot sensibly affirm (2.3a) and the negation of (2.3b) at the same time. To put this in more general terms, if we have propositions p and q such that p ⇒ q, we know that the complex proposition ‘p & not-q’ is necessarily false. (2.3a) has other entailments, as shown in (2.4). (2.4) a.  Moira has arrived in Edinburgh ⇒ Moira is not in Birmingham b. Moira has arrived in Edinburgh ⇒ Moira was travelling to Edinburgh The word arrived is an important contributor to (2.3a) having the entailments shown. If lived or been were substituted for arrived, the entailments would be different. If someone were to ask what arrive meant, a sentence like (2.3a) could be given as an example, explaining that it means that Moira travelled from somewhere else and is now in Edinburgh. However, the entailments from a sentence depend not only upon the content words in the sentence, like arrive, but also upon their grammatical organisation. For instance, the grammatical construction with has, sometimes called the “present perfect” construction, is crucial to the entailment in (2.3c) – if we replaced has with had, this entailment would no longer be valid, as Moira may subsequently have left Edinburgh again. There is more detailed discussion of the present perfect construction, alongside other ways of expressing tense and aspect in English, in Chapter 6.



sense rel ations 19

If (2.3a) is understood and accepted as true, then none of its entailments need to be put into words. They follow automatically, and can be inferred from (2.3a) by the hearer. It is obviously crucial to successful language use for speakers to make sure that the sentences that they utter have the correct entailments. In Chapter 1 I introduced the idea of sense as the aspects of the meaning of an expression that give it the denotation it has. We can think of the sense of a word in terms of the particular entailments that a sentence has as a result of containing that word. Whichever aspects of the word’s meaning are responsible for the sentences having those entailments are its sense. We can think of the entailment in (2.3c) as part of the sense of arrived, for instance: it is crucial to the meaning of arrived that if someone “has arrived” in a place, it follows that they are now “in” that place. 2.2 Compositionality

Given the potentially infinite supply of distinct sentences in a language, semanticists aim to develop a compositional theory of meaning. The principle of compositionality is the idea that the meaning of a complex expression is determined by the meanings of its parts and how those parts are put together. The idea that human language is compositional in this sense has a very long history, and is of great importance to linguistics: among other things, it offers a partial explanation of how we can comprehend the meanings of infinitely many different potential sentences, just by knowing the meanings of finitely many different sentence parts (and their combining rules). The meaningful parts of a sentence are clauses, phrases and words, and the meaningful parts of words are morphemes. The idea that the meaning of a complex expression depends on the meaning of the parts and how they are put together is hopefully a reasonably intuitive one. It’s comparable to what happens in arithmetic. Several things affect the result of an arithmetical operation, as we see in (2.5): it makes a difference what numbers are involved (2.5a), what operations are performed (2.5b), and – where there are multiple operations – also the order in which the operations take place (2.5c). (2.5) a. 3 + 2 = 5 but 3 + 4 = 7 b. 3 + 2 = 5 but 3 × 2 = 6 c. 3 × (2 + 4) = 18 but (3 × 2) + 4 = 10 The linguistic examples in (2.6) show something similar to what we see in (2.5c). Here we are considering a word that consists of three morphemes, un, lock and able. Our operations are not addition and

20

an introduction to english semantics and pr agmatics

­ ultiplication, but negation (or reversal) performed by the prefix unm and the formation of “capability” adjectives by the suffix -able. (2.6) a. un(lockable) ‘not able to be locked’ b. (unlock)able ‘able to be unlocked’ As in (2.5c), the brackets in (2.6) indicate the scope of the operations: that is to say, which parts of the representation un- and -able operate on. In (2.6a), un- operates on lockable, whereas ‑able operates only on lock. In (2.6b), un- operates only on lock and -able operates on unlock. However, unlike in arithmetic, there isn’t a default order of precedence, and we don’t use brackets like this in ordinary writing or speech. In practice, unlockable is just ambiguous – and the same goes for various other expressions which take un- as a prefix and -able as a suffix, such as unbendable, undoable and unstickable. Of course, sometimes only one interpretation makes sense: as it is not usually possible to unbreak something, unbreakable must be understood as ‘not able to be broken’. In syntax too there can be differences in meaning depending on the order in which operations apply. We saw an example of this in the first chapter, repeated here as (2.7); (2.8) is another straightforward example. (2.7) Jane saw the guy with binoculars. (2.8) I didn’t sleep for two days. Both of these are ambiguous for essentially the same reason as unlockable: we can think of them as involving two possible “bracketings”. On one reading of (2.7), with binoculars can be bracketed together with the guy as a constituent of the sentence: on the other reading, with binoculars modifies saw. On one reading of (2.8), for two days is an adjunct to sleep and the sentence expresses the meaning that ‘For two days, the speaker did not sleep’; on the other reading, for two days is a complement to sleep and can be bracketed with it, and the sentence expresses the meaning ‘It is not the case that the speaker slept for two days’. As with unlockable, the ambiguity is not a one-off fact about these particular sentences, but occurs systematically with similar sentences. (2.9a–c) are additional examples in which, like in (2.8), a prepositional phrase could be an adjunct or a complement, giving rise to different meanings. Our account of how compositionality works is going to have to allow for the fact that, in cases like this, sentences systematically end up with multiple possible interpretations. (2.9) a. I won’t be in town until 4 o’clock. b. I refuse to see him twice a day. c. I agreed to contact her during the committee meeting.



sense rel ations 21

At the same time, our language also contains expressions which appear to be compositional but in fact are not. These are expressions for which the meanings cannot be worked out by considering the meanings of the parts and how those parts fit together. We call them idioms. Classic examples in English are expressions such as head over heels, see eye to eye or kick the bucket. These simply have to be learned as wholes (see Grant and Bauer 2004 for more discussion). In this respect, idioms behave like individual words that consist of one morpheme: all we can do is to learn the association between the form and its meaning. As we have seen, for some words that consist of multiple morphemes, like unlockable, we can work out the meaning(s) of the word from the meanings of the individual morphemes and how they are combined. But there are also words that appear to consist of multiple morphemes but are idiomatic in their overall meaning: we can’t predict the meaning of greengrocer just from knowing the meanings of green and grocer, or that of greenhouse from knowing the meanings of green and house. 2.3 Synonymy

Synonymy is equivalence of sense. The nouns mother, mom and mum are synonyms of each other. When a single word in a sentence is replaced by a synonym – a word equivalent in sense – then the literal meaning of the sentence is not changed: My mother/mom/mum is from London. Sociolinguistic differences – things like the fact that mum and mom are informal, and that they are used respectively in British English and North American English – are not relevant to the sense of the word here, because they do not affect the propositional content of the sentence (although sociolinguistics is a huge and important area of study in its own right, which we won’t attempt to get into in this book). Sentences with the same meaning are called paraphrases of one another. (2.10a) and (2.10b) are paraphrases, differing only by substitution of the synonyms impudent and cheeky. (2.10) a. Andy is impudent. b. Andy is cheeky. c. Andy is impudent ⇒ Andy is cheeky d. Andy is cheeky ⇒ Andy is impudent e. *Andy is impudent but he isn’t cheeky. f. *Andy is cheeky but he isn’t impudent. As indicated by (2.10c), (2.10a) entails (2.10b) – again, assuming that we are talking about the same person called Andy at the same point in time. Similarly, (2.10b) entails (2.10a). If one of these entailments failed,

22

an introduction to english semantics and pr agmatics

(2.10a) and (2.10b) would not be paraphrases of each other. Also, both (2.10e) and (2.10f) are contradictions: it’s not possible for someone to be cheeky if they are not impudent, or vice versa. This also follows from (2.10a) and (2.10b) being paraphrases. In order to produce a semantic description of English, we would typically start from judgements about sentences and try to use those to establish sense relations between words. If we find a pair of sentences such as (2.10a) and (2.10b) which contain a different adjective but are otherwise identical, and it transpires that each entails the other (that is, they are paraphrases), then that would be evidence that the two adjectives are synonyms. Similarly, if we find a pair of sentences such as (2.10e) and (2.10f) that are both contradictions, that would be evidence that the two adjectives contained in those sentences are synonyms: if they were not perfectly synonymous, at least one of these sentences could at least potentially be true. The relation of paraphrase depends upon entailment: paraphrase means that there is a two-way entailment between the sentences. We can think of this as entailments indicating the sense relations between words, and sense relations indicating the entailment potentials of words. How can we find paraphrases? We might do this by observing language in use; but we might also invent test sentences and see whether or not particular entailments are present, according to our own judgement as language users (or the judgement of other proficient speakers of the language). To make this task easier, we might be interested in developing examples about which it is easy to have intuitions, such as (2.11). (2.11) a. You said Andy is cheeky, so that means he is impudent. b. You said Andy is impudent, so that means he is cheeky. So generally signals that an inference is being made. If both (2.11a) and (2.11b) are judged to be true, it appears that the entailments (2.10c) and (2.10d) both hold, and hence cheeky and impudent are synonyms. Note that we are interested in accessing our knowledge about the general pattern of entailment, not about the likely character of any specific person named Andy. We don’t want to know that if Andy is impudent then he is probably cheeky, or vice versa: we want to know that, if a speaker is committed to the idea that Andy is impudent, they are automatically committed to the idea that he is cheeky, and vice versa. If people accept (2.11a) and (2.11b) as reasonable arguments, they must agree with this, and we can conclude that the adjectives are synonyms. Having said that, we might need to allow for the possibility that someone will reject (2.11a) or (2.11b) – or both – on the basis that there is a difference in register between cheeky and impudent. That is to say,



sense rel ations 23

the circumstances under which you would use cheeky, as a speaker, are not the same as the circumstances under which you would use impudent. Like the contrast between mum and mother, this is part of our knowledge of language, but part of our sociolinguistic knowledge rather than our semantic knowledge. So one of the challenges we have to confront when we are trying to have intuitions about entailment is whether we, as judges, are relying only on the kinds of knowledge that we are interested in as analysts. Other pairs of synonymous adjectives include silent and noiseless, brave and courageous, and polite and courteous. But there are also many pairs of adjectives that have similar meanings without being synonyms. Consider (2.12). (2.12) a. The building is enormous. b. The building is big. Here, (2.12a) entails (2.12b), but the reverse is not true: the building could be big without qualifying as enormous. Hence, the relation between big and enormous is not one of synonymy. Synonymy not only applies to nouns and adjectives: it is also present in other word classes. The adverbs quickly and fast are synonyms, and, in Scottish English, so are the prepositions outside and outwith. And, as shown in the mother/mom/mum example, synonymy is not restricted to pairs of words: as another example, the triplet sofa, settee and couch are all synonymous. We can tell this because each pair of words within the triplet exhibits synonymy. In fact, because of the way entailment works, synonymy is a transitive relation: that is to say, if a is a synonym of b and b is a synonym of c, then a must also be a synonym of c. 2.4  Complementarity, antonymy, converseness and incompatibility

Some pairs of adjectives, such as moving and stationary, not only apply to a broad class of objects but also divide that whole class of objects into two non-overlapping sets. Everything that is capable of moving or being stationary – that is to say, any physical object – is, at a given point in time, either moving or stationary. Some other adjectives that divide their relevant domains in this way are listed in (2.13). (2.13) same different right wrong true false intact damaged connected disconnected

24

an introduction to english semantics and pr agmatics

These pairs are complementary terms – so called because the complement of things that are not described by one term are described by the other. That is to say, they give rise to the pattern of entailments illustrated in (2.14). (2.14) a. Maude’s is the same as yours. b. Maude’s is different from yours. c. Maude’s is the same as yours ⇒ Maude’s is not different from yours d. Maude’s is different from yours ⇒ Maude’s is not the same as yours e. Maude’s is not the same as yours ⇒ Maude’s is different from yours f. Maude’s is not different from yours ⇒ Maude’s is the same as yours In other words – assuming that Maude’s and yours denote the same thing in (2.14a) and (2.14b) – when (2.14a) is true, (2.14b) is false, and vice versa. This is a way of expressing the obvious idea that if two things are the same, they are not different; and that if two things are different, they are not the same. Another, more formal, way of putting this is that (2.14a) entails and is entailed by the negation of (2.14b), and (2.14b) entails and is entailed by the negation of (2.14a). Hence, (2.14a) is a paraphrase of the negation of (2.14b), and (2.14b) is a paraphrase of the negation of (2.14a). So we can think of complementaries as negative synonyms. Note that we have had to change the sentence frame slightly between (2.14a) and (2.14b), so it might be more precise to say that there’s complementarity between the same as and different from (rather than just same and different). A similar sense relation, but slightly weaker than complementarity, is that of antonymy. Sometimes this term is used just to mean any kind of “oppositeness”, but most semanticists use it to apply to one particular kind of oppositeness, exemplified by the adjectives noisy and silent, as in (2.15). Note that we use the symbol ⇏ to mean that an entailment is not valid: that is, if the sentence before ⇏ is true, it doesn’t necessarily follow that the sentence after ⇏ is true. (2.15)

a. b. c. d. e. f.

The street was noisy. The street was silent. The street was noisy ⇒ The street was not silent The street was silent ⇒ The street was not noisy The street was not noisy ⇏ The street was silent The street was not silent ⇏ The street was noisy



sense rel ations 25

Again assuming that we are referring to the same thing by the street in (2.15a) and (2.15b), we get the pattern of entailments shown above, which is different to the complementary case. If we know that (2.15a) is true then we know that (2.15b) is false, and if we know that (2.15b) is true then we know that (2.15a) is false. However, unlike the complementary case, knowing that (2.15a) is false doesn’t tell us whether or not (2.15b) is true, and knowing that (2.15b) is false doesn’t tell us whether or not (2.15a) is true. This is because there is middle ground between silent and noisy, whereas there is no middle ground between the same as and different from: to say that something is not noisy is not to say that it is silent, and to say that it is not silent is not to say that it is noisy. To put it another way, pairs of antonyms typically tap into meanings that are at opposite extremes, but unlike complementaries, they leave gaps in the middle. Under this definition, there are many pairs of antonyms: happy and sad, full and empty, early and late, and so on. It is not a coincidence that it is easier to find pairs of antonyms than it is to find synonyms or complementaries. Synonyms can be thought of as something of a luxury: given that two synonyms (such as courteous and polite) give rise to the same entailments, we could really do without one of them in the language, and still manage to convey all the information that we need to. Having an additional term might enable us to communicate in a more expressive or sociolinguistically richer style. Having words for both members of a complementary pair is arguably something of a luxury too: we could get away with having just one, and use negation to convey the other (instead of false we could just say not true). However, this will not work with antonyms: to say that something is full is more than just saying that it is not empty. We need both full and empty in the language in order to talk about quantity in this way. A general feature of the adjectives that form antonym pairs is that there is also a sense relation between their comparative forms. Comparatives are formed by the suffix -­ er for some adjectives (thicker, poorer, humbler) or more generally by the construction “more + adjective” (more patient, more obstinate). The comparative forms of an antonym pair of adjectives exhibit a sense relation called converseness, illustrated in (2.16). (2.16) a. France is bigger than Germany. b. Germany is smaller than France. c. France is bigger than Germany ⇒ Germany is smaller than France d. Germany is smaller than France ⇒ France is bigger than Germany

26

an introduction to english semantics and pr agmatics

Starting with (2.16a), if we replace bigger with smaller and swap the position of the noun phrases France and Germany, we obtain a paraphrase, (2.16b). Thus we say that bigger and smaller are converses. We can think of converseness as something like a version of synonymy that also requires the reordering of noun phrases. Converseness is also present in other word classes, including nouns (such as parent (of) and child (of)), verbs (such as precede and follow) and prepositions (such as above and below). In each of these cases, if we have two entities X and Y, and X stands in one of these relations to Y, it must be the case that Y stands in the converse relationship to X (for instance, if X is above Y, then Y is below X, and vice versa). Just as synonymy is not restricted to pairs of items, neither is ­antonymy. We can often identify sets of terms for which any two members are antonyms: we can say that these sets exhibit incompatibility. For instance, we can consider a set of colour adjectives such as {black, blue, green, yellow, red, white, grey} to be incompatible, in that if we say something is “blue”, it follows automatically that it is not black, green, yellow, etc. – assuming that we are dealing with objects with a single predominant colour. Similarly, a set of nouns denoting shapes, such as {triangle, circle, square}, might also exhibit incompatibility: if something is “a triangle”, it is not a circle or a square, and so on. A set has incompatibility if every member of the set exhibits antonymy with every other member of the set, so the diagnostics for incompatibility will be essentially the same as for antonymy. 2.5 Hyponymy

The relation of hyponymy is about the different subcategories of a word’s denotation. The pattern of entailment that defines hyponymy is illustrated in (2.17). (2.17) a. There’s a house on the riverbank. b. There’s a building on the riverbank. c. There’s a house on the riverbank ⇒ There’s a building on the riverbank d. There’s a building on the riverbank ⇏ There’s a house on the riverbank If it is true that there is a house on the riverbank, it follows that there is a building on the riverbank, as indicated in (2.17c). This is because a house is one kind of building. There are other kinds of building: school, church, factory, and so on. Hence, the reverse entailment does not hold, as shown in (2.17d): the building on the riverbank could be something other than



sense rel ations 27

a house. This pattern of entailment tells us that we are dealing with hyponymy, and that house is a hyponym of building, or equivalently building is a superordinate (sometimes called “hypernym”) of house. Often, as in (2.17), a sentence containing a hyponym entails the corresponding sentence in which the hyponym has been replaced by its superordinate. However, if the sentence is negative, this pattern may be reversed, as shown in (2.18). (2.18) a. There isn’t a house on the riverbank. b. There isn’t a building on the riverbank. c. There isn’t a house on the riverbank ⇏ There isn’t a building on the riverbank d. There isn’t a building on the riverbank ⇒ There isn’t a house on the riverbank To put it another way, the fact of there being a building on the riverbank is a necessary condition for there to be a house on the riverbank. Hence, it would be reasonable to say that ‘building’ is a component of the meaning of house: a house is a ‘building for living in’. 2.5.1  Hierarchies of hyponyms

House is a hyponym of the superordinate building, but building is in turn a hyponym of structure, and structure is in turn a hyponym of thing. Like synonymy, hyponymy is also transitive: if a is a hyponym of b and b is a hyponym of c, then a is a hyponym of c. This means that house is also a hyponym of structure and a hyponym of thing, which is fairly obvious if we think of the definition above: if we replaced building by structure or thing, the entailment patterns in (2.17) and (2.18) would stay the same. Similarly, thing is a superordinate of house, and so on. These relations are illustrated in Figure 2.1. Incidentally, this means that we don’t have to worry about whether we’ve captured all the layers in this hierarchy. If we discovered that there was another category between structure and thing, say artefact, which was a hyponym of thing and a superordinate of structure, that would not affect the hyponymy relation between structure and thing. As we move up a hierarchy of hyponymy, the senses of the words become less specific and their denotations become larger and more general. At lower levels, the senses are more detailed and the words denote narrower ranges of things, as illustrated in Figure 2.2. In Figure 2.2 we try to capture the meaning of each hyponym as the meaning of its immediate superordinate with a modifier (e.g. ‘for living in’). This captures the key idea that a hyponym is a special case of its

28

an introduction to english semantics and pr agmatics

thing

superordinate of structure, building, house, . . .

structure

hyponym of thing, superordinate of building, house, . . .

building

hyponym of structure, thing, superordinate of house, . . .

house

hyponym of building, structure, thing, superordinate of . . .

Figure 2.1 An example of levels of hyponymy thing

‘physical entity’

structure

‘thing with connections between its parts’ = ‘physical entity with connections between its parts’

building

‘structure with walls and a roof’ = ‘physical entity with connections between its parts with walls and a roof’

house

‘building for living in’ = ‘physical entity with connections between its parts with walls and a roof, for living in’

Figure 2.2 Hyponym senses get successively more detailed as we go down the tree

superordinate: it’s an instance of the superordinate that has certain properties that not all instances of that superordinate share. In practice, it’s not always easy to provide useful hyponym definitions in this way, because sometimes we run into circularity: a dog is a type of animal, but it’s difficult to describe what type of animal it is without using the word ‘dog’ or some related word like ‘canine’. It is, in effect, ‘an animal that is a dog’. Even so, hyponym hierarchies are useful to us as language users because they guarantee the validity of a large number of inferences. If someone tells us facts about a particular house, we know that the things they are telling us are also true of at least one structure, and at least one building, and so on. To take a marginally more practical example, if we know that platypus is a hyponym of mammal, we know that (2.19a) entails (2.19b). (2.19) a. Platypuses lay eggs. b. There are mammals that lay eggs.



sense rel ations 29

entity

thing

time

place

structure product

plant

dam

building vehicle

barn

house

person1

animal1

tool animal2

garden tool

idea

student

person2 freshman

(post) graduate

kitchen utensil saw hacksaw

Figure 2.3 Part of the hyponym hierarchy of English nouns

Hyponym hierarchies exist for other parts of speech, such as verbs and adjectives, as well as nouns. Amble is a hyponym of walk, which in turn is a hyponym of move; (made of) oak is a hyponym of wooden, and so on. Moreover, these hierarchies are potentially vast. WordNet is a systematic database of English word meanings, which, at the time of writing, contains entries for more than 155,000 words. In creating WordNet, Miller and Fellbaum (1991: 204) discovered that a hyponym hierarchy with twenty-six high-level superordinates (time, plant, animal, and so on) ‘provides a place for every English noun’. Figure 2.3 shows a tiny fraction of the WordNet noun hierarchy, featuring just seven of the twenty-six superordinates (and, of course, omitting the vast majority of their hyponyms). Note that some entities appear twice in this hierarchy: we distinguish two senses of person (corresponding roughly to ‘human’ and ‘psychological individual’) and two senses of animal (corresponding roughly to ‘living thing that is not a plant’ and ‘living thing that is not a plant or a human’). In addition to hyponymy itself, we might try to identify other sense relations within a hyponym hierarchy. The hyponyms of a given superordinate might be linked by incompatibility – house and factory are different hyponyms of building, so we might argue that something that is a house is not a factory – but actually this does not follow from

30

an introduction to english semantics and pr agmatics

the ­definition of hyponymy we have been working with here. Indeed, if house and dwelling are synonyms, then they are both hyponyms of building, but they are clearly not incompatible. Summary

In this chapter, we discussed the notion of entailment and its relevance to semantics. Based on this, we were able to define various distinct sense relations that can obtain between expressions in a language: synonymy, antonymy, complementarity, converseness, incompatibility and hyponymy. Entailments between sentences are the evidence for sense relations between words; or, going the other way, sense relations indicate the entailment potentials of words. The idea of sense relations is helpful to us, as analysts, when trying to give a semantic description of a language, but also when we as speakers are trying to learn the semantics of a language. Exercises

1. The word dishonest means ‘not honest’. The following five words also have ‘not’ as part of their meaning: distrust, disregard, disprove, dislike, dissuade. Write a brief gloss for the meaning of each, similar to the one given for dishonest. Thinking of sentences for the words will probably help. Use the term scope (introduced in Section 2.2) to describe the difference between the glosses. 2. Here is an unsatisfactory attempt to explain the meaning of not good enough: not good means ‘bad or average’; enough means ‘sufficiently’; so not good enough means ‘sufficiently bad or average’. With the aid of brackets, explain why the phrase actually means ‘inadequate’. 3. Someone once said to me, “You and I are well suited. We don’t like the same things.” The context indicated (and I checked by asking) that the speaker meant to convey that we are well suited because the things we don’t like are the same: but the sentence is in principle ambiguous. Explain the ambiguity, and comment on unambiguous alternatives. 4. Which of the following sentences entail which? (a) The students liked the course. (b) The students loved the course. (c) The rain stopped. (d) The rain ceased.



sense rel ations 31

5. Provide example sentences and write out a pattern of entailments (comparable to (2.10a–d)) that establishes soundless, silent and noiseless as synonymous. 6. Are awake and asleep complementaries? Give reasons for your answer. Whether you have answered yes or no, how would you include halfawake, half-asleep and dozy in an account of the meanings of awake and asleep? 7. The hyponyms of footwear include shoes, sneakers, trainers, sandals, slippers, boots and galoshes. Draw up a hyponym hierarchy for these words. Is footwear the superordinate that you use for all of the hyponyms, or are there other alternatives? Recommendations for reading

Cruse (2011) provides a thorough discussion of oppositeness in meaning, as well as hyponymy. Lappin (2001) provides a good article-length introduction to formal semantics, and Saeed (2015) complements this by dealing in greater detail with some relevant topics. WordNet is available and browsable online at https://wordnet.princeton.edu/.

3 Nouns

Overview

Nouns form the majority of English words. They typically denote entities with rich and complex sets of properties. We can think of some of these properties as being associated with hyponymy relations, as discussed in Section 2.5: because a dog is a type of mammal, we know that dogs possess all the properties that mammals must have. But additional sense relations apply to nouns, chief among them the has-relation, which we discuss in this chapter. The has-relation captures the fact that the things denoted by nouns can have parts, whether these are physical or conceptual, and the question of which parts a noun has may be highly relevant to its meaning. This chapter also discusses the distinction between count and mass nouns: that is, between nouns that are treated as denoting entities that can be separated and distinguished from one other, and nouns that are not. 3.1 The has-relation

The everyday words square, circle and triangle are also technical terms in geometry, where they have tight definitions. We might define a square as a ‘closed figure which has four straight sides of equal length separated by 90° angles’. The fact that a square has four sides is built into its definition. We can think of the entity a square as being associated with ‘four sides’ by the has-relation. Like the relations discussed in Chapter 2, the has-relation is important to semantics because it guarantees the truth of certain entailments, as illustrated in (3.1). (3.1) a.  That figure is a square ⇒ ‘That figure has four sides of equal length’ b.  That figure is a square ⇒ ‘That figure has four internal 90° angles’ 32



nouns 33

Mathematical terms are somewhat atypical of natural language because they have such unambiguous definitions. Trying to define nouns in everyday use – part of the task of linguistic semantics – isn’t always straightforward. In particular, the status of has-relations is sometimes unclear. Consider (3.2). (3.2) a.  Mary drew a face ⇒ ‘The picture that Mary drew includes eyes’ b.  Tom drew a house ⇒ ‘The picture that Tom drew includes a door’ If we think of the things that are denoted by the English word face, the examples that spring to mind probably include eyes, a nose and a mouth, among other features. That is to say, if we think of a prototype of a face – a clear, central example of the denotation of face – it probably has eyes, a nose and a mouth. Similarly, a prototype of a house probably has a door, windows, a roof, and so on. (Conversely, there are numerous features that many real faces and houses have but which are not likely to be present in the corresponding prototype – say, a goatee, or a carport, respectively.) The inferences in (3.2) are based on the relations ‘a face has eyes’ and ‘a house has a door’. But these has-relations are really only valid if we think in terms of prototypes. Something could be unambiguously a face without having eyes, and unambiguously a house without having a door. If Mary drew a picture of someone wearing shades, we would grant that “Mary drew a face”, but the inference in (3.2) wouldn’t be valid. In (3.1), then, we’re dealing with proper entailments of the kind ­introduced in Section 2.1. But in (3.2) we have weakened the entailments by making them conditional on prototypicality. It would be more appropriate to write them down as something more like (3.3). (3.3) a.  Mary drew a face ⇒ ‘If the picture that Mary drew is prototypical, it includes eyes’ b.  Tom drew a house ⇒ ‘If the picture that Tom drew is prototypical, it includes a door’ Essentially, it’s necessary to weaken things in this way because typical English words are not as tightly defined as technical words like square (in the geometric sense). But it is also useful for us to do this because knowledge about prototypical properties is an important part of our knowledge of what words mean. In fact, for many nouns there are very few properties that are completely essential to the definition, even though a noun may have many prototypical properties. This point has been argued by a number

34

an introduction to english semantics and pr agmatics

of influential thinkers about language, so I want to dwell on it for a moment. We can say that a square must have four sides, and anything which does not have four sides cannot correctly be called a square. But can we say anything similar about a face, or a house, or many other things? The classic example in this respect is the word game. Wittgenstein (1953) argued that there were no features at all that were shared by all the things we call a game. Not all games involve competition, or skill, or physical ability; not all games have rules; not all games have playing pieces or a scoring system. So, by this token, knowing that something “is a game” doesn’t tell you anything about the has-relations that it possesses. Similarly, although we can think of a prototypical house as having windows, a door and a roof, it would not automatically cease to be a house if it had no windows, or lost its roof. Despite this lack of obligatory properties, it still seems perfectly reasonable to identify something as being a game or not being a game, or to talk about a building being a house or not being a house. Consequently, a useful idea is that word meanings are organised (at least in part) around prototypes rather than obligatory properties – an idea called “prototype theory”, which is particularly associated with Eleanor Rosch and colleagues. Calling something a game does not guarantee that it possesses any one specific property associated with the prototypical game, but it does suggest that that thing possesses enough of the prototypical game properties, to a sufficient extent, to fall within the category. Similarly, calling something a house doesn’t guarantee that it has windows, a door or a roof, but it does strongly suggest (in the absence of indications to the contrary) that it has most or all of these features, along with other prototypical house features. In the following sections we will look at some of the consequences of the has-relation as applied to various aspects of noun meaning. In doing this, we’ll generally be making some tacit assumptions about prototypicality when talking about word meanings, although we will also see some of the limitations of this approach, for instance in the discussion of the has-relation and hyponymy in Section 3.1.2. 3.1.1  Inferring existence from the has-relation

By appealing to the has-relation, we can infer the existence of entities that haven’t been explicitly mentioned. Consider (3.4). (3.4) Some kids walked up to a house, knocked on the front door and ran away.



nouns 35

In (3.4), we have an indefinite article, a, and a definite article, the. A noun phrase that first introduces its referent into conversation is usually indefinite, whereas subsequent mentions of the same referent will usually involve a definite noun phrase. This is why (3.5a) is reasonably natural but (3.5b) is odd (assuming that a house denotes the same house in both sentences) – (3.5b) attempts to refer to the already-established referent with an indefinite noun phrase. (3.5) a. Some kids walked up to a house. The house was old and spooky. b. Some kids walked up to a house. *A house was old and spooky. What we see in (3.4) is that, after mentioning a house, front door behaves as though it has also already been mentioned: it’s acceptable to say the front door, and it would be odd to say a front door. To put it another way, if we say a house and then say the front door, the hearer is able to infer that we probably mean ‘the front door of the just-mentioned house’. Clark (1975) introduced the term bridging inference to describe this kind of inference, as it involves connecting up the newly mentioned material to that which has been mentioned before. The pattern shown in (3.5) holds to some extent for non-prototypical features. Earlier I mentioned that carport was a non-prototypical feature that a house might have. In a context like (3.5), we can still use the with carport, but it may also be fine to use a. That doesn’t work with front door, as we see in (3.6). (The situation with door is a little more complicated because a door might suggest that the house has multiple doors.) (3.6) a. Some kids walked up to a house. The front door was to the right of them. b. Some kids walked up to a house. ?A front door was to the right of them. c. Some kids walked up to a house. The carport was to the right of them. d. Some kids walked up to a house. A carport was to the right of them. These examples show that we are conscious of the has-relations that are associated with the nouns we mention, and these can influence how we talk about things that we subsequently mention. Prototypical parts may require the use of definite articles, whereas parts that are not prototypical can be used with indefinite articles. To put it another way, the hearer can reasonably infer from the speaker’s use of a house that the door of that house exists, and they expect to encounter the

36

an introduction to english semantics and pr agmatics

expression the door if the speaker intends to refer to that door. The hearer may also be willing to accommodate the use of an expression like the carport if they are willing to draw the inference that the house has a carport, but as this is not part of the prototype, using a carport is also fine. Chapter 10 will go into a little more detail about how the use of definite and indefinite articles feeds into our understanding of discourse. 3.1.2  Hyponymy, prototypes and the has-relation

The has-relation is obviously not quite the same thing as hyponymy, discussed in Section 2.5, but these two relations interact in important ways. Recall that hyponymy is about categories being grouped under superordinate terms. To take another geometrical example, we could say that square is a hyponym of quadrilateral, and so are rectangle, parallelogram, kite, rhombus and trapezium. Hyponyms then “inherit” the parts that their superordinates have (Miller and Fellbaum 1991: 206). By definition, a quadrilateral has exactly four sides – or, to put it another way, it is connected to the attribute “exactly four sides” by the has-relation. This same relation is inherited by square, rectangle, parallelogram, kite, rhombus and trapezium. Quadrilateral is in turn a hyponym of polygon: by definition, a polygon has straight sides. This relation, “has straight sides”, is inherited by all the hyponyms of polygon, including quadrilateral and all its hyponyms. This kind of inheritance is important to our semantic knowledge. As a result of it, when we learn about the hyponymy relations that a word enters into, we automatically acquire knowledge about its has-relations. However, although this idea is easy to apply to terms with clear definitions, such as mathematical shapes, it becomes more complicated when we are dealing with prototypes. The prototype of a hyponym does not generally inherit all the has-relations from the prototype of its superordinate. For instance, the Neolithic houses uncovered at Skara Brae in Orkney had no windows: if we coin the neologism skara for a house that resembles one of these houses, skara will be a hyponym of house, but the prototypical skara will have no windows. This observation applies not only to has-relations but also to other properties. A classic example is that a property of a prototypical bird is that it “can fly”. This property is inherited by most of the hyponyms of bird, but of course not all: penguin is a hyponym of bird and the prototypical penguin cannot fly. We can’t even say with confidence that the properties of a prototypical superordinate will be inherited by most of its hyponyms. Suppose we had 10,000 words for different kinds of penguin.



nouns 37

They would all be hyponyms of bird, but it wouldn’t alter the fact that a prototypical bird can fly and a penguin cannot. A hyponym will, of course, inherit all the obligatory properties, including has-relations, from its superordinate(s). But, as we already discussed in the case of game, some seemingly well-understood superordinate categories don’t seem to have any obligatory properties – so in these cases knowing about a hyponym–superordinate relationship (for instance, knowing that oware is a game) doesn’t necessarily tell us anything extra about the properties that the hyponym must have. Still, it might be reasonable to think that prototypical properties of superordinates are somewhat likely to be inherited by at least some of the (prototypical instances of) their hyponyms. 3.1.3  Parts can have parts

Just as hyponyms can themselves have hyponyms, so the parts of an object – the things that that object has – can themselves have parts. We could say, for instance, that a suburb has houses, a house has windows, a window has panes, and so on. As far as the properties of these parts are concerned, remember that again we’re dealing with the has-relation rather than hyponymy. Even if properties are obligatory for a whole, they need not be obligatory of its parts. A square obligatorily has four sides and four corners, but it is not true that “each side has four corners” or “each corner has four sides”. 3.1.4  Spatial parts

For practical reasons, we’re often interested in talking about physical objects which don’t necessarily have separate distinguishable parts. Even so, we might want to talk about aspects of the physical nature of these objects. We can do this by using words like top, bottom, sides, front, back, and so on. Although not all physical objects have these particular spatial parts, many different kinds of object do – heads, buses, screens, wardrobes and pianos, to name but a few. Some objects are associated with a particular physical orientation – a bus falls into this category. There is no doubt about which end of the bus is the front and which end is the back: these are intrinsic to the nature of the thing. And arguably the same is true of top and bottom – even if a bus is upside-down, there is a real sense in which the top of the bus still refers to its roof. By contrast, other objects are not associated with a particular orientation, or only partially associated with one. A tree has a top but doesn’t have a front, for instance; and a rock doesn’t usually have any of

38

an introduction to english semantics and pr agmatics

these things (assuming that the issue of which way up it happens to be is a geological accident rather than a matter of design). The lack of intrinsic orientation doesn’t mean that we can’t talk about the front of a rock or the back of a tree. But it does mean that the use of these terms depends on the context of utterance: that is to say, these expressions are deictic. Like the examples of deixis we saw in Section 1.2.1, the meaning of front or back depends on who is speaking and, in this case, specifically how they themselves are located. A reasonable guess is that what the speaker means by “the front of the rock” is the part nearest the speaker, and what they mean by “the back of the tree” is the side of the trunk that is furthest away from them. In practice, things are often even more complicated than this. One complication is that, although speakers usually use expressions like front or back relative to where they are located, they may choose to use them relative to where the hearer is located. A second complication is that speakers sometimes use expressions like front or back relative to their own position, or that of the hearer, even for objects which do have an intrinsic orientation. Suppose that there is a chair in the middle of the room that is pointing in a northerly direction (that is, if you were sitting normally on it, you would be facing north). Suppose that you are standing two metres to the south of the chair, and I am standing two metres to the west of the chair, and we’re both facing towards the chair. Now I say “Please place a coin in front of the chair”. (This is very unlikely in real life, but psycholinguistics experiments are full of bizarre instructions like that.) There are at least three ways in which you could interpret that request. Taking in front of to refer to the intrinsic orientation of the chair, I’m asking you to place a coin to the north of the chair. Taking it to be deictic based on the speaker’s location, I’m asking you to place a coin to the west of the chair. Taking it to be deictic based on the hearer’s location, I’m asking you to place a coin to the south of the chair. Essentially, in common with a lot of deictic expressions, many references to the spatial parts and orientations of objects are technically ambiguous. But, in common with a lot of technically ambiguous expressions in language, much of the time we seem to use them perfectly well without misunderstanding one another. The question of how we do this is still an open one and continues to animate a lot of research. 3.1.5  Ends and beginnings

Long, thin objects, such as those listed in (3.7a), have ends. Sometimes things are sufficiently orientable that it makes sense to distinguish two



nouns 39

different kinds of end, which we could broadly term beginnings and ends, although for some objects we use more technical vocabulary to describe these (a ship has a bow at one end and a stern at the other, a river has a source at one end and a mouth at the other). Nouns denoting periods of time, such as those listed in (3.7b), also have beginnings and ends, as well as middles. By extension, we can think of events and processes that take place over time as having beginnings, middles and ends, as in the examples in (3.7c). In Chapter 6, we will see some of the ways in which the structure of events influences how we have to talk about them, from a grammatical standpoint. (3.7) a. rope, string, train, ship, road, river, canal b. day, week, month, era, term, semester, century c.  conversation, demonstration, ceremony, meal, reception, match 3.1.6  Body part terms, metaphor and has-relations

Words for body parts are widely used in English (and many other languages) to describe many different kinds of object and their parts: the foot of a mountain, the neck of a bottle, the mouth of a river, and so on. These uses started out as metaphors, presumably because of some general human tendency to interpret the world in terms of our own bodies, although in many cases these have long since become established as distinct senses of the relevant words (sometimes referred to as “dead metaphors”). There is more detailed discussion of figurative language, including metaphor, in Chapter 9. Here I just want to make one point about how has-relations work in some of these cases. Compare (3.8a), (3.8b) and (3.8c). (3.8) a. [I] hired a forest ranger on his weekend off, and started climbing one of the faces of Mount Rushmore. (, retrieved 18 April 2022) b.  Between the mountains Machu Picchu and Huayna Picchu form a face looking at the sky. (, retrieved 18 April 2022) c.  Hulking perilously above the village of Grindelwald in Switzerland is the infamous north face of the Eiger. (, retrieved 18 April 2022)

40

an introduction to english semantics and pr agmatics

In (3.8a), face means ‘a face’ – well, more precisely, ‘a carving depicting a face’. It is not metaphorical and does not refer to the geographical feature of that name. We can expect that the face in (3.8a) inherits the has-relations of a prototypical face: we expect the carving to depict eyes, a mouth, and so on. By contrast, in (3.8b), face means something more like ‘rock formation that happens to resemble a face’. This is more like a metaphor: we understand the speaker not to be claiming that these peaks literally constitute a face, but rather to be pointing out the resemblance. Here we might expect that some of the parts of a face will be somehow discernible, but perhaps not all of them – just enough to make the comparison plausible. And in (3.8c), face has a purely “dead metaphor” status. We don’t expect the Eiger to display anything recognisable as the parts of a face: the word face is used simply because that is what the side of a mountain is called. In short, the has-relations of body part terms are applicable when we are using the terms literally, inapplicable when we are dealing with dead metaphors, and potentially partially applicable – but not to be relied on – when we are dealing with new “living” metaphors. And there is nothing special about body part terms in this respect; we can expect to observe the same pattern for other expressions that can be used metaphorically. 3.2  Count nouns and mass nouns

Another striking feature of noun meanings is that we can distinguish two broad categories, which are represented differently in the grammar of English: count nouns, such as loaf and coin, and mass nouns, such as bread and money. As shown in Table 3.1, there are a lot of differences between count and mass nouns in terms of how they combine with expressions of quantity, as well as the singular indefinite article a. This reflects the obvious fact that count nouns denote distinguishable whole entities whereas mass nouns denote undifferentiated, and hence uncountable, substances. As the examples in Table 3.1 illustrate, the division between count and mass nouns cross-cuts other semantic differences – the words loaf and bread enable us to talk about the same thing in a ‘count’ or ‘mass’ way. There are also words such as cake which can be freely used as either count or mass nouns. As the table suggests, there are also cases where what is usually a mass noun can be “coerced” into a count use – if someone talks about many breads, you might infer that they mean ‘many distinct varieties of bread’. There are also cases where what is usually a count noun can be coerced into a mass interpretation – if someone says



nouns 41

Table 3.1  Distinguishing between count and mass nouns Count nouns

Mass nouns

This is a loaf. This is a coin. How many loaves are there? How many coins are there? a large number of loaves a large number of coins six loaves six coins ?some loaf *some coin some loaves some coins *How much loaves are there? *How much coins are there?

?This is a bread. *This is a money. ?How many breads are there? *How many monies are there? ?a large number of breads *a large number of monies ?six breads *six monies some bread some money ?some breads ?some monies How much bread is there? How much money is there?

there is egg all over the floor, we can easily interpret that as referring to the contents of broken eggs rather than one or more unbroken individual eggs. In essence, where we have both count nouns and mass nouns for the same thing, the speaker has a choice about how they are going to portray reality: as containing individual objects like loaves or coins, or as containing undifferentiated “stuff” like bread or money. Of course, this does not mean that when a speaker uses a mass noun they become incapable of distinguishing its parts: when we talk about furniture we can still tell the difference between tables and chairs (and between individual chairs), but we choose not to emphasise that point. Hyponymy (and incompatibility) exists among mass nouns just as among count nouns: velvet, corduroy, denim and so on are incompatible hyponyms of the mass noun cloth, and are themselves mass nouns. In principle, we might expect mass nouns not to enter into has-relations, because homogeneous substance is not separable into distinct parts: but this is perhaps not a clear-cut issue, simply because (as remarked above) whether we treat something as mass or count doesn’t necessarily say a great deal about its physical reality. If we agree that cloth has threads, then this is a has-relation involving a mass noun, and it is inherited by the hyponyms of cloth in the usual fashion.

42

an introduction to english semantics and pr agmatics

Summary

In this chapter we have considered some of the characteristic properties of nouns, with particular emphasis on the has-relation and the interplay between this and other sense relations. We have seen that the has-relation is a potentially powerful tool for learning about the entailments of hyponyms of nouns, but this is sometimes a complicated matter because has-relations are often not obligatory, and prototypical properties are not always inherited by hyponyms in the same way as obligatory properties. We have also considered the spatial properties of objects and some of the ways in which language allows us to talk about these properties. And we considered the distinction between count nouns and mass nouns as a way of portraying the world: labelling with a mass noun treats the thing referred to as a homogeneous substance without distinct parts, and this has consequences for how we can talk about this referent. Rather like the choice of how we talk about spatial configuration, the distinction between mass nouns and count nouns does not correspond to a crisp distinction between two kinds of things that exist in the world, but rather between two different ways in which we can relate to the things in the world through language. We will encounter similar ideas in the following chapters, which deal in turn with adjectives and verbs. Exercises

1. What parts does a prototype shoe have? Do those parts have parts? 2. If we wanted to describe the meanings of some spatial part words, we might say something like this: “The top of a thing is one of its sides, the side that is uppermost. The bottom of a thing is one of its sides, the side that is down. The front is one of the sides, the side that faces forwards. The back is one of its sides, the side that faces away from the front.” If this description is correct, what sense relations hold between the words side, top, bottom, front and back? Give reasons to support your answer. 3. Paper, glass and cheese are ambiguous between a count sense and a mass sense. (a) Devise a pair of example sentences for each of them that clearly brings out the count–mass difference.



nouns 43

(b) Find some hyponyms for each of the words in each of its senses. Use these to comment on the systematic difference in meaning between the count and mass interpretations of these words. 4. In the question “Have you ever eaten rabbit?”, what difference does the lack of an article (rabbit instead of a rabbit) make to the interpretation of the noun rabbit? 5. Why might we interpret left differently when we are describing something as being to the left of the chair versus to the left of the stool? Recommendations for reading

For prototype theory, Mervis and Rosch (1981) give a detailed overview of some of the major ideas and their consequences. Johannsen and De Ruiter (2013) and Coventry et al. (2018) report useful work on how speakers choose spatially referring expressions; Weisberg and Chatterjee (2020) specifically examine how ultimate Frisbee players do this. Kearns (2011) is an excellent source for more about differences in meaning between mass and count nouns.

4 Adjectives

Overview

In Chapter 2 we already talked about a number of the sense relations that adjectives (along with other parts of speech) can enter into: synonymy, complementarity, antonymy, incompatibility and converseness. In this chapter, we focus on two more specific aspects of adjective meaning – their tendency to express different degrees, or levels, of the quantity that they denote, and the way they combine with noun meanings. In both cases we will also see how contextual factors influence the interpretation of adjectives. 4.1 Gradability

In Section 2.4, we talked about several sense relations that apply to adjectives. One of them was converseness, as illustrated by (4.1). (4.1) a. France is bigger than Germany. b. Germany is smaller than France. Converseness relies on the existence of comparative forms like bigger and smaller, and these comparatives are possible because the adjectives they are based on are gradable. That is to say, the language allows us to express different degrees or levels to which nouns possess the qualities that these adjectives denote. Some adverbs are also gradable, as we see with quickly in (4.2a), but nouns and verbs generally are not. To make any sense of (4.2b) we have to interpret fish as fishlike (an adjective), and to make sense of (4.2c) we have to interpret more swims as swims more, which isn’t quite the same thing. (4.2) a. A salmon swims more quickly than a human. b. *A salmon is more fish than a human. c. *A salmon more swims than a human. 44



adjectives 45

The use of the comparative -er suffix in (4.1) is an indicator that we are dealing with a gradable property. So is the use of more in (4.2a), as well as the use of than in both (4.1) and (4.2a). Other examples of morphemes that indicate the presence of gradability are shown, underlined, in (4.3). (4.3) a. He is the rudest person I have ever met. b. These questions are too difficult. c. How long is this going to take? d. The weather was very hot. e. That is a good enough set of examples. The adjectives in the examples in (4.3) are all members of antonym pairs: rude–polite, difficult–easy, long–short, hot–cold, good–bad. Each adjective denotes what we can think of as a region towards one or the other end of a scale. On the scale of “difficulty”, difficult denotes values towards the high end and easy denotes values towards the low end. Difficult and easy are not complementaries, in the terms introduced in Section 2.4, because some things are neither difficult nor easy on the “difficulty” scale but somewhere in between. Curiously, these pairs of antonyms don’t behave quite the same way within the language system. Consider (4.4). (4.4)

a. b. c. d.

This will take a long time. This will take a short time. How long is this going to take? *How short is this going to take?

In this particular context, if we want to ask a question about where we are on the scale of “duration”, we can only use long – the option of short seems ungrammatical. More often we have a situation in which one member of the antonym pair is used to ask neutral questions and the other is not, as in (4.5). (4.5) a. How old is that player? b. How young is that player? (4.5b) would be an odd question to ask about a professional footballer who is, say, thirty years old. It might be a reasonable question in a context in which the player in question is or appears to be unusually young, and it might be a reasonable question if the player ought to be young (for instance, because it is a junior tournament) but doesn’t appear to be. But (4.5a) is acceptable in any of these situations. In fact, what seems to be happening here is that old is the preferred adjective for talking about the “age” scale (as in I am thirty-eight years old) and if a speaker chooses to use young instead, they invite the hearer to infer that

46

an introduction to english semantics and pr agmatics

there must be some kind of special reason for their choice. In Section 8.2.4, we will discuss an account of pragmatic inference that proposes a more general explanation of how meanings like this arise when a speaker uses an unusual or disfavoured expression. 4.1.1  Non-gradable adjectives

Not all adjectives are gradable. In particular, the members of complementary pairs are sometimes resistant to being interpreted this way. In Section 2.4, I gave six examples of complementary pairs: moving–­stationary, same–different, right–wrong, true–false, intact–damaged and c­ onnected–disconnected. All these pairs, by definition, partition all the entities that they apply to into two, non-overlapping classes: everything that is moving is not stationary, and vice versa. However, while things that are moving move at varying speeds, we can’t really say that something that is moving faster is “more moving” than something else, and we can’t say that two stationary objects differ in “how stationary” they are. So the moving–stationary pair separates objects into two categories but then doesn’t really discriminate within those categories at all. By contrast, complementary pairs like same–different do not follow this pattern. In this case, if two things are the same, that’s absolute – we can’t quantify “how same” they are. But different is gradable: chalk and cheese are more different than apples and pears. In short, complementary pairs vary quite considerably in how gradable they are, as shown by the examples in (4.6). (4.6)

a. ?The jury heard evidence that was too false to be credible. b. ?The jury heard evidence that was too true to disbelieve. c. ?That is the rightest thing any politician has said on the topic. d. ?That is the wrongest thing any politician has said on the topic. e. ?How intact is the vase? f. How damaged is the vase? g. I feel very connected to the community. h. I feel very disconnected from the community.

Other kinds of adjectives can also be thought of as non-gradable. In (4.3a), we had an example of a superlative, rudest. Superlatives denote extreme ends of scales, and this makes them resistant to being further graded: their position on the scale is not up for negotiation, so we have no need to clarify it further. In the case of *bestest or *more fastest, we could think of this as just reflecting a morphological or syntactic rule that stops us combining a superlative or comparative form with another



adjectives 47

superlative or comparative form. But actually we see a similar pattern with other adjectives that denote extreme values, as in (4.7): these are also difficult to use gradably (and when we try, it might just be understood as a form of emphasis that doesn’t really change the meaning, as in (4.7c)). (4.7) a. *The presentation was very excellent. b. ?The weather was too freezing. c. ?She is extremely unique. Conversely, there are some things that can be done with these extreme adjectives that can’t be done with ordinary gradable adjectives. For instance, they can be modified with “maximiser” adverbs such as absolutely and completely, whereas gradable adjectives cannot, as shown in (4.8). (4.8)

a. b. c. d. e. f.

*The shrink wrapping was absolutely thin. *Her performance was completely good. You’ll look completely different with your hair restyled. What you say is absolutely true. Digital sound reproduction is completely perfect. The weather has been absolutely freezing.

Even when extreme adjectives aren’t gradable, they can still influence how we talk about gradable quantities. For instance, if a speaker has the possibility of describing something as excellent, that may influence the circumstances under which they will describe that thing as merely good – and hearers are aware of this. The possibility of drawing additional inferences based on the speaker’s failure to use a (relatively) extreme description turns out to be very useful in communication. There will be more on that in Section 8.2.2. 4.2  Combining adjective meanings with noun meanings

When an adjective modifies a noun, how do the meanings of the adjective and the noun combine? This seems like a rather trivial matter, and in some cases it is – but there are also cases where it is surprisingly complicated. In this section I will just introduce a couple of cases that illustrate that idea. Let’s dispose of the relatively trivial case first. If we consider an example like green bicycles, we could say that this adjective–noun combination denotes all the things that are both green and bicycles. If we use a Venn diagram to represent the denotations of green (things) and bicycles, as in Figure 4.1, then exactly those things that fall into the intersection of the two ellipses are green bicycles.

48

an introduction to english semantics and pr agmatics

green things

green bicycles

bicycles

Figure 4.1 The simplest case of an adjective modifying a noun is like the intersection of sets

This seems to correspond reasonably well with our intuitions about what green bicycles means. So, for cases like this, we can think of the combination of an adjective and noun meaning as simply involving the intersection of two sets – the set of things denoted by the noun and the set of things to which the adjective applies. For that reason, we call adjectives like this intersective adjectives. Things get more interesting, and challenging, when we consider adjectives that are not intersective. In the following subsections we’ll briefly consider three kinds of non-intersective adjectives: those which end up with potentially broader denotations than we would expect on intersective grounds; those which end up with a denotation that doesn’t overlap with the intersection at all; and those which are crucially dependent on context for their interpretation. 4.2.1  Non-intersective adjectives with broader denotations

One slightly unexpected category of non-intersective adjectives ­comprises words like alleged, assumed, supposed, and so on. When these combine with a noun, the resulting phrase has a denotation that ­potentially includes entities that fall within the usual denotation of the noun as well as entities that do not. Consider the underlined phrases in (4.9). (4.9) a.  Turkish authorities have cracked down on alleged supporters of the U.S.-based Muslim cleric Fethullah Gülen following a failed coup in 2016. (, retrieved 20 April 2022) b. After months of boring assumed winners, the Oscars race has been shaken up recently by shocker developments no one saw coming. (, retrieved 20 April 2022)





adjectives 49

c.  They lament the failure to prosecute any of the supposed masterminds behind the insurrection. (, retrieved 20 April 2022)

The speaker of (4.9a) is clearly signalling the possibility that not all the people cracked down on are actual supporters of Gülen. In (4.9b), the speaker is talking about people (and movies) that are or have been assumed to achieve the status of winners in the future. And in (4.9c), the speaker seems to be expressing scepticism towards the idea that any of the people involved could really be called masterminds. In general, then, this class of adjectives is clearly non-intersective. What we get when we compose alleged and supporter is not ‘a special kind of supporter, specifically one that also has the property of being alleged’. Rather, we get a meaning along the lines of ‘someone who has been alleged (by someone else) to be a supporter’. A similar pattern holds for assumed and supposed. All these examples are compositional, in the sense that the meanings are predictable from the meanings of the adjectives and the meanings of the nouns: it’s just that the denotation of alleged supporter is not a subset of the denotation of supporter. The three examples above all involve adjectives that are derived from verbs, and all of them result in a meaning that implicitly involves the action of that verb (someone alleging, assuming or supposing something). But there are also non-intersective adjectives that are not directly based on verbs, as we see in (4.10). (4.10) a. Miller continued to lead, looking a likely winner for 16 laps. (, retrieved 20 April 2022) b. If Tiger is at 2020 levels, then by that standard he would be a possible winner but a long shot. (, retrieved 20 April 2022) c. Giannis Antetokounmpo may be playing like an MVP, but he’s an unlikely winner at this point. (, retrieved 20 April 2022) In these examples, none of likely winner, possible winner or unlikely winner is a specific kind of winner. In each case, what the adjective contributes

50

an introduction to english semantics and pr agmatics

is some kind of evaluation of the probability of the individual being talked about actually going on to be a winner of the relevant kind. The class of possible winners for a given event is likely to be much bigger than the class of winners. So, once again, these are non-intersective. We use the word modality to describe the kind of meaning that these particular adjectives express, and it will be discussed in more detail in Section 7.1. 4.2.2  Privative adjectives

In the previous subsection we talked about the adjective unlikely, which seems to suggest that the entity being talked about probably doesn’t belong to the denotation of the following noun. However, we don’t seem to be able to use impossible in quite the same way, as shown in (4.11) – perhaps if we are using the word impossible, we are not licensed to call the person a winner in the same breath. (4.11) *Tiger is an impossible winner at this point. Surprisingly, though, there are adjectives which do explicitly rule out the possibility that the entity in question belongs to the usual denotation of the following noun. These are sometimes called privative adjectives, and the most widely discussed example is fake, as in (4.12). (4.12) World’s costliest painting Salvator Mundi is a fake Leonardo da Vinci, claims documentary. (, retrieved 20 April 2022) Here the use of fake Leonardo da Vinci explicitly excludes the possibility that the painting is due to Leonardo da Vinci. The denotation of Leonardo da Vinci here includes all the works by Leonardo, rather than the individual himself (that is to say, this is an example of the figure of speech called metonymy, which is discussed a little more in Section 9.3). So, ostensibly, the denotation of fake Leonardo da Vinci doesn’t even overlap with the denotation of Leonardo da Vinci. Of course, the noun still contributes meaning to the expression, even in this privative case. A fake Leonardo is not the same thing as a fake Picasso or a fake Warhol. If you were to paint a version of Salvator Mundi, it would be odd to describe the result as a fake Leonardo, unless you made some special effort to pass it off as Leonardo’s work. So in this case the noun is contributing the meaning of what the work is purporting to be, and the adjective is contributing the meaning that it isn’t a genuine example.



adjectives 51

There is a possible analysis on which this kind of adjective isn’t really privative at all, though: that would be if we take Leonardo not only to denote all the works by Leonardo himself, but also those that closely resemble those works or are claimed to be genuine examples. Then we can separate the category of Leonardos into fake Leonardos and genuine Leonardos. This analysis is not particularly intuitive as applied to fake, but perhaps makes sense if we think about words like former, as in (4.13). (4.13) Barack Obama is a former President of the United States. Here, we could think of former as privative if we understand that Obama is not the President of the United States, and cannot be President again (because of term limits); or we could think of it as simply intersective if we take President of the United States to denote anyone who has held that office at any point in time. 4.2.3  Relative adjective meanings

In Section 4.2, we discussed green as an example of an intersective adjective. Green, in its sense as a colour word (rather than its sense of ‘environmentally conscious’ or that of ‘naïve’, among others), is reasonably stable in meaning from context to context, although what counts as green may vary from noun to noun (must the object be completely of that colour, or only partially?). However, a lot of adjectives have meanings which are much more context-dependent than this, as in (4.14). (4.14) a. Sucrose is a big molecule. b. The swimming pool was the size of a small lake. If we wanted to describe big or small as intersective, we would have to say that the denotation of big molecule is anything that is both a member of the set of ‘big objects’ and a member of the set of ‘molecules’, and that the denotation of small lake is anything that is both a member of the set of ‘small objects’ and a member of the set of ‘lakes’. But this is obviously wrong – it’s hard to imagine that the category of ‘big objects’ contains any individual molecules, or that the category of ‘small objects’ contains any lakes. Moreover, it seems intuitively obvious that what the speaker of (4.14b) means is ‘the swimming pool was big’ – so clearly the same thing can be small according to one standard and big according to another standard. What seems to be happening here is that we are interpreting the adjectives big and small in a way that is relative to the noun they are modifying. These adjectives (and many others, such as wide, narrow, long, short, and so on) contribute a meaning that restricts the denotation of

52

an introduction to english semantics and pr agmatics

the noun to examples that possess the relevant property to a sufficient degree. These are sometimes described as relative intersective adjectives. They are intersective only in the weaker sense that the denotation of the adjective–noun combination is a subset of the denotation of the noun alone: all big molecules are molecules, all small lakes are lakes, and so on. But, in order to figure out precisely what subset we are talking about, we have to consider the (perhaps rather idiosyncratic) way in which the specific adjective combines with the specific noun. Clearly this is possible – we use relative intersective adjectives all the time without seeming to be misunderstood very often – but the question of how exactly it works is quite a difficult one. There are also plenty of adjectives that have different senses which attract different kinds of interpretation. Consider (4.15). (4.15) She’s a good politician. As applied to people, good has a possible sense in which it means something like ‘morally upstanding’; but in the context of (4.15), it also has a possible sense in which it means something like ‘effective at doing the job’. If what we mean by (4.15) is that the person being described is both a politician and a morally upstanding person, then we are using good as an intersective adjective. If we mean that the person is effective in their job as a politician, then we are using good as a relative intersective adjective. In practice, we can sometimes exclude the possibility of fully intersective interpretations by appeal to common sense: if I describe something as a good shovel, you can assume that I’m not trying to ascribe moral qualities to an inanimate object, and for that reason I must mean that it’s an effective implement for use in the act of shovelling. However, we may sometimes need more sophisticated strategies than this for resolving potential ambiguities of this kind. Summary

This chapter introduced some additional features of adjective meaning. The characteristic of gradability, which applies to some adjectives, relates to their ability to convey different levels or strengths of a property, and has some interesting consequences for how we can and cannot use and further modify those adjectives. We can also distinguish adjectives by how their meanings combine with those of the nouns they modify. In some cases, this takes place in the obvious way, but in other cases the interaction of meanings is surprisingly complex. Indeed, a single adjective may have multiple senses which modify the noun in dif-



adjectives 53

ferent ways, creating ambiguity. We are able to deal with some of these ambiguities, however, by appealing to context and world knowledge, a theme which will recur throughout this book. Exercises

1. The adverb quite has two different meanings when it modifies adjectives. In one sense it is a “downtoner”: quite friendly can be glossed as ‘moderately friendly’. In another sense it is a “maximiser”: quite exceptional is synonymous to ‘exceptional to the fullest extent’. More specifically, quite is a downtoner with words such as clever, late, small and unusual, but a maximiser with right, finished, impossible and alone. What is the relevant difference in types of meaning between these classes of words? 2. Which of the following adjectives would normally yield biased questions if you inserted them into the frame “How [adjective] is/was x?”, and which would normally yield unbiased questions? old young rude polite unpalatable tasty weak strong 3. Which of the following phrases can be explained as examples of intersective adjective use, and which cannot? Why? royal visitor royal correspondent heavy eater wise fool 4. Proxima Centauri is a small, cool, red star located near our Solar System. In fact, it is the closest other star to the Sun. Comment on the meaning of the adjectives in the context of these two sentences. Recommendations for reading

The relevant sections of general grammars of English such as Quirk et al. (1985) and Huddleston and Pullum (2002) are rich sources for detailed accounts of adjective meaning. Kennedy (2012) provides a very readable discussion of adjective use and meaning, and some of its broader philosophical implications. There are many more complex

54

an introduction to english semantics and pr agmatics

papers on the semantics of gradable adjectives (for instance, McNally 2011), which get into some very interesting questions about the nature of the underlying “scale of measurement” for these adjectives, but these papers tend to assume a lot more prior knowledge.

5 Verbs

Overview

This chapter is about verb meanings. One way of thinking about what makes up a sentence such as Robby brought me the news is that the verb is at the heart of it and it “says something about” the referents of the noun phrases that surround it (here Robby, me and the news). More specifically, we could think of the verb as explaining how these noun phrases relate to one another. Although we can think of other parts of speech in a similar way, it is customary to consider the verb as the semantic centre of the sentence (or clause). In this chapter we’ll consider how different verbs impose different requirements on the rest of the clause, specifically in terms of how many other referents must be introduced, and we’ll look at how the verbs cause different roles to be associated with those referents. 5.1  Verb types and arguments

One way of categorising verbs is according to how many referring expressions they interrelate: that is to say, how many noun phrases (NPs) they require. Usually they require one, two or three, as illustrated by the examples in (5.1), where the NPs are underlined. (5.1) a. Billy sings. (one noun phrase) b. Ella admires Beethoven. (two noun phrases) c. I offered her a scone. (three noun phrases) In practice, verbs do not necessarily require NPs in all these positions. Some verbs will accept preposition phrases (PPs) in certain positions, such as to her in (5.2a). And sometimes a position can be filled by a more complex kind of structure, as in (5.2b–c). (5.2) a. I offered a scone to her. b. The report confirms that spring came early. 55

56



an introduction to english semantics and pr agmatics

c.  That the daffodils are blooming confirms that spring came early.

In (5.2b) and (5.2c), positions are filled by embedded clauses introduced by that. A clause is a structure that usually has a verb of its own and can carry a proposition on its own: for example, spring came early contains the verb came and expresses a proposition about the start of the season. However, in (5.2b–c), this clause is not free-standing but has been embedded into another clause as the object of the verb confirm. (5.2c) also contains another that-clause embedded in subject position. The word that is one of the markers made available by English grammar to mark a clause as embedded. We use the term argument to cover all the kinds of obligatory constituents that verbs require, whether they are NPs (the report), PPs (to her) or embedded clauses (that the daffodils are blooming). (5.2a) has three arguments; the main clauses in (5.2b–c) each have two arguments. A verb that requires both a subject argument and a direct object argument, such as admires in (5.1b), is called transitive. We could add an additional argument to (5.1a) – for example, in the bath – but the additional argument is not necessary for the sentence to be grammatical. (5.1a) is therefore intransitive. Other verbs, such as offer in (5.1c), obligatorily require both a direct object and an indirect object: we cannot normally say *I offered her or *I offered a scone (unless the potential recipient is contextually obvious). (5.1c) is therefore ditransitive. In (5.2a), we see a slightly different version of this in which the indirect object is replaced by a PP, with a corresponding change in word order. It is sometimes useful to distinguish one further class of sentences, copular sentences such as John is my brother (or That’s what I’m talking about, as in Chapter 1), where the verb is is used to “predicate properties of a subject” (that is, label the subject as having particular properties). We can delve a bit further into the different categories of verbs. Intransitives have been divided into two rather opaquely named kinds (Trask 1993: 290–2) on the basis of the type of subject argument that the verbs require, as follows: • An unergative verb requires a subject that is consciously responsible for what happens. Walk is such a verb, and tourists walk through the eco park is an unergative clause. A good test for unergativity is acceptability with the adverb carefully, because taking care is only a possibility when an action is carried out deliberately. • An unaccusative verb is an intransitive verb in which the subject is affected by the action but does not count as being responsible for it.



verbs 57

Grow, drop and die are verbs of this kind. These verbs do not occur readily with adverbs such as carefully: *Mort carefully died. From the hearer’s point of view, knowing that a verb is unergative potentially tells you something about the subject, namely that it is something capable of conscious action. Of course, this is usually rather trivial: in the example above, we already know that tourists are capable of conscious action. However, we can also distinguish categories of verbs which give rise to richer and more elaborate kinds of entailment, as we shall see in Section 5.2. 5.1.1  Other kinds of arguments

In a way, the notion of arguments is more general than just ‘the obligatory constituents that verbs require’. I mentioned in the introduction to this section that we could think of other parts of speech as fulfilling similar functions to verbs, and I meant that specifically in terms of how they relate different arguments. Consider (5.3). (5.3) a. Maya sent Bill the letter. b. The letter moved from Maya to Bill. We would typically analyse (5.3a) as a ditransitive sentence with the verb sent, the subject Maya, the direct object the letter and the indirect object Bill. But (5.3b) relates the same arguments in quite a similar way, even though the verb moved is intransitive and both the arguments from Maya and to Bill are optional. And abstractly there is nothing to stop us from saying that (5.3b) is a sentence in which from Maya has two obligatory arguments, a verb (expressing the nature of the motion from Maya) and an NP (identifying the thing that moves), plus an optional argument (the destination). Sometimes it can be useful to think about other parts of a sentence or clause in terms of the arguments that are needed to combine with them in order to make a complete sentence. However, usually we focus on the verb and its arguments, and that will be the case in the rest of this chapter. 5.2  Causative verbs

In Section 2.2, we talked about the principle of compositionality: the idea that the meaning of a complex expression is determined by the meaning of its parts and the way those parts are put together. Focusing on verbs, we can think of the meaning of a clause as being determined not only by the meaning of the verb itself but also by the kinds of

58

an introduction to english semantics and pr agmatics

a­ rguments that it has. The arguments can in fact reshape our impression of the meaning that the verb contributes, as we see in (5.4). (5.4) a. Robbers spray victims to sleep. (Fiji Post, 1 June 1995) b. Yet Birthday’s reckless spontaneity has been focus-grouped to death. (, retrieved 22 April 2022) In the case of (5.4a), we would normally understand spray as meaning something like ‘distribute a substance in aerosol form’. We would expect the clause to have an obligatory subject, as it does – robbers – and also a direct object, identifying what the substance was sprayed at or onto, which it does – victims. We might also see the substance itself being identified, which it is not in (5.4a). Instead, we have this unexpected argument to sleep. To make sense of this, we have to assume a different interpretation for spray, one in which it is causative, because the reading of to sleep as indicating the effect of the spraying is the only way to make sense of its presence here as an argument. Similarly, in (5.4b), we might expect the verb focus-group (meaning something like ‘expose something to a focus group’) to have two arguments, a subject and a direct object: here it occurs as a passive, so it has no explicit subject (for more on passives, see Section 10.3.3). But it still has two arguments, the direct object Birthday (denoting a song with that title) and to death. Again, this suggests a causative interpretation of focus-group, with the speaker of (5.4b) understood to be saying that, as a consequence of being exposed to a focus group, the song Birthday has – in some metaphorical sense – died. The meaning expressed by a causative sentence is, in general, that a situation is caused by whatever the subject noun phrase refers to, and the situation is described by the additional argument. In (5.4a), that situation is The victims sleep, and in (5.4b) it is Birthday has died. The sentences in the left-hand column of Table 5.1 are further examples of causatives, with each one entailing the sentence to its right. The sentences on the right in Table 5.1 describe states or events, without making reference to their causes. The causatives on the left share the following properties: • They include a causative verb: that is to say, one that expresses causation. In these examples, we have the causative verbs make, get, force, cause, have, and prevent. Prevent is a negative causative. • The subject of the causative sentence is used to refer to whatever entity – concrete or abstract – brings about the situation described by the sentence on the right.



verbs 59

Table 5.1  Examples of causative sentences with an entailment from each Causatives

Entailments

The thought made her gleeful. The children got the kite to fly. Bad weather forces us to cancel the picnic. His inexperience is causing the decisions to go unactioned. I had the students study this article. The lock prevented him from opening the door.

She was gleeful. The kite flew. We are cancelling the picnic. The decisions are going unactioned. The students studied this article. He did not open the door.

• The causative has an argument which somehow carries the same proposition as the sentence to its right. The sentence affirms that that proposition is brought about by some cause. The verb cause itself is arguably a superordinate for the other causative verbs in Table 5.1: it expresses the vaguest and most general sense of causation. For example, we can think of force as a hyponym of cause that means more specifically something like ‘cause an unwanted or resisted consequence’. In fact, as we will see in Section 8.2.4, using cause in place of a more specific verb can sometimes turn out to have interesting ­consequences, in terms of the meaning that the speaker is taken to convey. Semantically, causative verbs have a minimum of two arguments, one denoting the causer and one denoting the caused state or event. We can call this latter argument the embedded situation. The embedded situation itself contains arguments. For the causative sentence I made John give Mary the book, the embedded situation ‘John gave Mary the book’ has three arguments, John, Mary and the book. 5.2.1  Identifying embedded situations

The exact way in which the embedded situation is expressed syntactically in a causative sentence is beyond the scope of this book. We might think of things like her gleeful or the kite to fly as being special kinds of clauses that express the embedded situation, but it’s controversial as to whether we can call them clauses, because they lack tensed verbs of their own. But for our purposes it’s enough to say that these expressions enable us to recover the embedded situations, such as ‘she was gleeful’ or ‘the kite flew’. In this latter case, the proposition receives its past tense from the verb got in the causative sentence. In other cases, embedded situations are even less clearly signposted. We can think of certain sentences as involving “understood” embedded

60

an introduction to english semantics and pr agmatics

situations, even when these are not made at all explicit. (5.5) is a potential example of this. (5.5) The staff nurse gave Lucinda a key for the week. As Tenny (2000) points out, for the week in (5.5) doesn’t modify the verb give – the speaker of (5.5) is not saying that the action of ‘the staff nurse giving Lucinda a key’ lasted the whole week. Instead, what for the week modifies appears to be an understood embedded situation: that is, the situation in which ‘Lucinda has a key’. That is to say, (5.5) is a concise way of expressing the meaning that ‘the staff nurse gave Lucinda a key, with the intention of causing Lucinda to have the key for the duration of the week’. There is some additional evidence that embedded situations which are not syntactically visible might figure in language use and interpretation. Consider the situation described in (5.6). (5.6) a. When the company started, their main selling point was good value. b. They then quietly doubled the price of their product. c. After a public outcry, the company lowered the price again. The use of again would usually be appropriate in a case where something had happened before – technically, it is often said to presuppose that the thing in question happened before. (There is more on presuppositions in Section 8.4.) But, in (5.6c), again appears to refer to the event in which ‘the company lowered the price’, even though that didn’t happen before. This use of again is called restitutive: it involves the restoration of a previously existing state of affairs (Tenny 2000), in this case the one in which the company sells its product at a lower price. A good way of capturing this would be to say that again is actually referring to the embedded situation in which ‘the company sells its product at a lower price’. The existence of this situation seems to be part of the meaning of (5.6c), even though it is not explicitly expressed there. Table 5.2 shows some more examples of causatives that involve just a single clause, but which nevertheless entail a proposition about a caused situation (as shown in the right-hand column). We could think of these as involving various kinds of understood embedded situations, which in these cases are bound up in the verb semantics: although feed and eat are different verbs, an act of feeding naturally involves an embedded situation which contains eating, and so on. The last two lines of Table 5.2 show causatives entailing unaccusatives which use the same verb form: Gardeners grow vines ⇒ Vines grow; He broke a bone ⇒ A bone broke. When we have two different types of



verbs 61

Table 5.2  Three kinds of one-clause causative with an entailment from each Causatives

Entailments

different verbs (e.g. feed–eat) She fed the baby some mashed banana. The baby ate some mashed banana. The bank has lowered its interest rate. The bank’s interest rate dropped. Drought killed the lawn. The lawn died. morphologically related verbs and adjectives (e.g. enrich–rich) Nitrogen spills have enriched the soil The soil is rich here. here. The graphic artist enlarged the logo. The logo became larger. His job deafened Dougie. Dougie became deaf (to an extent). same verb form used causatively and non-­causatively (e.g. walk–walk) The guide walks tourists through the Tourists walk through the eco park. eco park. The gardener grew several vines. Several vines grew. He broke one of his bones. One of his bones broke.

e­ xpression with the same form, we sometimes call this conversion or zero derivation. Here, an unaccusative verb is formed by the conversion of a causative verb (or vice versa) without any change in form. (Conversion can also take place between other word classes: we saw an example earlier, with focus group being a verb derived from the existing noun.) Fellbaum (2000: 54) notes that many English verbs exhibit just this kind of causative-to-unaccusative entailment with conversion: some common examples are bend, dry, hang, hurt, lean, pop, spill, split and turn. However, although this is clearly a productive and regular pattern, there is nothing intrinsically special about this type of entailment: similar entailments can arise between lexically unrelated word forms, such as kill and die in Table 5.2. 5.3  Thematic relations

As we’ve already seen, verbs differ not only in how many arguments they require, but also in how those arguments are related to the verb semantically – whether they are the subject, the direct object, an embedded situation, and so on. This is clearly related to the syntactic structure of the sentence. However, we have also seen that sentences with different syntactic structures can assign the same relations to their arguments, as in (5.7).

62

an introduction to english semantics and pr agmatics

(5.7) a. I offered her a scone. b. I offered a scone to her. The term thematic relation, also called thematic role or participant role, is used to describe the role that the referent of a noun phrase plays in its sentence. In both (5.8a) and (5.8b), the relationship between I and offered is the same: the individual denoted by I is performing the action denoted by offered. Similarly, although the syntax is different, the relationship between her and offered is the same in both cases: the individual denoted by her is at the other end of the action denoted by offered, in that she is the prospective recipient. And in both cases a scone denotes the thing that is offered. Of course, the relationships I just described are not a special property of the two sentences in (5.7) but a generalisation about the verb offer in general: it requires three arguments, one of which performs the action of ‘offering’, one of which ‘receives the offer’, and one of which is ‘offered’. This is a more specific generalisation than just saying that offer is ditransitive: that is, that it takes three arguments. Also, it makes clear that offer taking three arguments is not an arbitrary fact about the verb, but rather it’s intrinsic to the verb’s meaning: sentences involving offer must have arguments that fulfil these vital thematic roles. This also gives us another way to think about the classification of verbs. Rather than just dividing them up according to how many arguments they require, we can group them according to the specific set of thematic relations that have to be fulfilled by the noun phrases that complete the sentence. In the case of offer, we might call these thematic relations Agent, Recipient and Theme – this last term describing an entity that undergoes an action without changing its state (here, the thing that is offered, in (5.7) a scone). We can then group offer with other verbs that also require an Agent, Recipient and Theme, such as give, donate, sell, pledge and (in at least one sense) promise. Various other thematic relations can be distinguished, and doing so might help us distinguish the meaning of verbs, even when they appear to have very similar syntactic structures, such as in (5.8). (5.8) a. Jane criticised Bill. b. The kettle scalded Bill. Here we could say that the subject of (5.8a), Jane, is deliberately performing an action, whereas the subject of (5.8b), the kettle, is not. We might want to distinguish these at the level of thematic relations by allowing that Jane is an Agent but the kettle is something else which we call a Stimulus. We could then distinguish those verbs like criti-



verbs 63

cise which require an Agent from those like scald which only require a Stimulus (although in this case could also occur with an Agent). We can also distinguish these from examples such as (5.9) in which the subject of the sentence does not perform or accomplish anything but is instead a recipient of some kind of input – in this case, we might call the subject, Bill, an Experiencer. (5.9) Bill heard the traffic. In a similar vein, we can distinguish Recipients and Themes from other classes of things that are acted upon in some way. In (5.10), the object of the sentence undergoes an action and with it a change of state, unlike a scone in (5.9), which remains unchanged: we could refer to the car in (5.10) as a Patient. (5.10) The tree damaged the car. And we can also distinguish the Recipient role from the subtly different role of Beneficiary: in (5.11), Valerie is presented as the person for whom an action has been performed, but (5.11) does not commit us to the view that Valerie actually gets to receive or experience the consequences of this action. (5.11) David wrote a poem for Valerie. Numerous other thematic relations have been proposed – Instrument, Location, Source, Goal, and so on. I won’t attempt to catalogue them all here. Theorists are interested in knowing whether there are just a relatively small number of possible distinct thematic relations, among which the verbs of English (or any other language) select their requirements – or whether there are lots of subtly different verb-specific thematic relations. For instance, was I justified in asserting that the person who is offered something is a Recipient, even if they don’t actually receive anything? Or should I categorise them as a Beneficiary, even if they don’t actually benefit? Or am I ultimately forced to say that they are an “Offeree”, a thematic relation that is unique to the verb offer? Similarly, we might say that the verb hit requires two thematic relations to be fulfilled, Agent and Patient, or we could argue that actually it specifies two much more specific thematic relations, “Hitter” and “Hittee”, and simplifying this to Agent and Patient would be invalid. If we can identify shared sets of thematic relations, this will be a useful way of identifying similarities in the semantics of broad classes of verbs – but if we can’t simplify in this way, thematic relations might not be a very useful idea for us. Different approaches to grammar make different claims about the connection between thematic relations and syntax. One idea of this

64

an introduction to english semantics and pr agmatics

kind, due to Chomsky (1981), is the “theta-criterion”: this claims there is a one-to-one correspondence between thematic relations (or their near-equivalent in the theoretical framework used in that paper, “thetaroles”) and syntactic arguments, and that any sentence in which there is a mismatch between syntactic arguments and theta-roles will not be well formed. We could think of examples such as (5.12) as illustrating this idea. In (5.12a), there are not enough noun phrases to satisfy the requirement that give imposes for there to be an Agent, a Recipient and a Theme. In (5.12b), there are too many arguments for the verb cry, which only requires an Experiencer. And in (5.12c), there is no possible Agent, as required by the verb request, because both the entities denoted by noun phrases in (5.12c) are inanimate. (5.12) a. *John gave to Mary. b. *Bill cried Jane. c. *The rock requested water. Summary

In this chapter, we have considered the role of verbs within a sentence. It can make sense to think of the verb as requiring a certain number of arguments, depending upon its semantics, in order for the sentence containing it to be semantically complete. We can classify verbs according to how many arguments they require, but we can also identify specific categories of verbs that cause their containing sentences to give rise to specific patterns of entailment. And we can drill further down into the issue of how to complete a sentence containing a verb, by noting that many verbs – on account of their meaning – require their arguments to have specific semantic properties. We can think of this as the verb specifying a set of thematic relations that must be fulfilled by the surrounding noun phrases in order for the sentence to be complete. On this view, each verb is associated with a complex set of requirements that it imposes upon its containing sentence – and this property of verb meaning is associated with the traditional idea of the verb as the semantic core of the sentence. Exercises

1. In February 2002 a UK government minister announced the resignation of a senior civil servant in his department. It was subsequently reported that the civil servant only found out about his own alleged resignation from listening to the radio. This led to a question being



verbs 65

asked by the media: Who is going to be resigned next? In fact, the civil servant subsequently resigned three months later. Resigning is supposed to be a conscious act performed by the person who quits the post. If, in talking about the situation described above, someone had used the expression The minister resigned the civil servant, would the sentence have been causative? Would it have the same meaning as The minister made the civil servant resign? 2. According to the nursery rhyme, after Humpty Dumpty’s accident, All the king’s horses / And all the king’s men / Couldn’t put Humpty together again. Given that no one had put Humpty together on any previous occasion, what kind of verb is put in this sentence, and how does its meaning relate to that of again? 3. What is each of the following sentences: unaccusative or unergative? Give reasons for your answers: (a) The kite flew. (b) My heart sank. (c) The students were reading. 4. We could think of pay as a hyponym of give. How does the argument structure of pay relate to that of give? What is the difference in argument terms between John paid Mary and John paid the bill? Recommendations for reading

Kearns (2011) and Huddleston and Pullum (2002) both offer illuminating accounts of the topics dealt with in this chapter. Lists of candidate semantic roles vary enormously between sources and between theories, an issue discussed by Croft (1991) and Dowty (1991). Dowty proposes to address this by positing more general Agent and Patient “protoroles”. Radford (2004) discusses Chomsky’s theta-criterion with reference to syntactic theory.

6  Tense and aspect

Overview

This chapter is about two important features of how English grammar allows us to convey meanings. First, we consider how we express when events occurred, especially relative to the time of speaking or writing, which is the role of tense. Secondly, we consider how we can convey by grammatical signals more about how an event occurred, with respect to time: that it to say, is it an ongoing event, a repeated event, or an event that took place at a single moment? This is the role of aspect. 6.1  Talking about events in time

In his novel The Restaurant at the End of the Universe, Douglas Adams claims that the major problem with time travel is that of grammar, and refers to an essential grammar book which, among other things, tells you ‘how to describe something that was about to happen to you in the past before you avoided it by time-jumping forward two days in order to avoid it’. Given that we don’t have time travel, our needs are somewhat simpler when it comes to describing events in time – but even then, this process can become surprisingly complex. Consider (6.1), from an article by Andrew O’Hagan (‘The end of British farming’, London Review of Books, vol. 23, 2001). (6.1) When I told people I was spending time with farmers, they’d say: how can you stand it, they just complain all day and they’ve always got their hand out. (6.1) is especially rich in time reference, so let’s take a moment to unpack what is going on here, and how time gets expressed and understood. First, we have the verb told, which is a past tense form of the verb tell. Specifically, it is the past simple form, so called because it is simple 66



tense and aspect 67

in aspect and does not involve any of the special aspectual meanings that will be introduced later in this chapter. It indicates that the first event described in (6.1), the speaker telling people that he was spending time with farmers, took place before the point in time at which (6.1) was uttered. So the tense marking on the verb expresses this time reference for the whole event, relative to another point in time. In this case, the time of the event is expressed relative to the time of utterance – and therefore the meaning that (6.1) expresses depends upon the time at which it is uttered. Hence, we can think of tense as deictic, much like the time adverbials (today, soon, and so on) we discussed briefly in Section 1.2.1. The next verb in (6.1), was spending, is in a form called past progressive. The tense marking on was (by contrast with am) indicates that this is a past form: English happens to mark tense on the first element in a multi-word verb form like this. The combination of the auxiliary be in front of the main verb (which could take the form be, am, is, are, was or were, as appropriate) with the -ing suffix on the verb itself indicates that we are dealing with progressive aspect, which portrays an event as being in progress at the relevant time point. Thus we infer that the event in which the speaker spent time with farmers was an ongoing one at the time the speaker told people about it. Next we have they’d say, which is a contracted version of they would say. Would is the past simple form of the modal auxiliary will (Chapter 7 has more on modal auxiliaries). There is no suffix in English that we can put on verbs as an indicator of futurity, so will is the main grammatical device for signalling future time reference in English. What we have here is the past form of a future marker: the speaker is taking us to some point in time prior to the time of utterance and looking ahead to a time when people say to him how can you stand it? In this case, the point in time the speaker is taking us back to is that same point in time at which he told people he was spending time with farmers: we can tell this because of the use of when right at the beginning of the sentence. The next verbs we come to are the verbs can . . . stand and complain. These are present simple. But, of course, in this case, that doesn’t mean that they refer to the time at which the speaker is uttering (6.1) – we’re still back at the past time in which these people are talking to the speaker about farmers. So, although we can use present simple forms to refer to events taking place at the time of utterance, we can also use them to refer to past or future events, if that is the current time reference point that we have set in the preceding context. In fact, they just complain all day is a little more complex still: it seems to convey not that ‘farmers complain all day at the present moment’ but rather ‘farmers

68

an introduction to english semantics and pr agmatics

habitually complain all day’. English does not have a grammatical marker of habitual aspect, but it can sometimes be inferred from the use of the present simple, and that seems to be an appropriate interpretation in this case. It might also be an appropriate interpretation of told – there are reasons to suppose that the speaker is not talking about a single event, but rather a pattern of events. There will be more on habituality in Section 6.3.1. Finally, we have the verbs have . . . got, in they’ve always got their hand out, which are in a form known as present perfect. This uses an idiom, to have one’s hand out, meaning roughly ‘to be asking for money or gifts’. Here it occurs in the present tense (we couldn’t say *They’ve got their hand out an hour ago, for instance) – and once again, that doesn’t mean that it necessarily refers to the time of utterance, but to the current time reference point, which in this case is in the past relative to when (6.1) is uttered. We can think of it as expressing the culmination of a past event: the farmers were ‘getting their hand out’ and have now done so. The use of always is best read as hyperbolic (see Section 9.6) and probably expresses something like habitual aspect – but for ­convenience here, let’s think of it as just marking an ongoing event more emphatically. In sum, (6.1) relates a whole bunch of events to each other in time. We can think of the events as follows: (a) ‘speaker utters (6.1)’ (b) ‘speaker tells people he is spending time with farmers’ (c) ‘speaker spends time with farmers’ (d) ‘people say “how can you stand it . . .”’ (e) ‘speaker can (or cannot) stand spending time with farmers’ (f) ‘farmers (are alleged to) complain all day’ (g) ‘farmers (are alleged to) get their hand out’ (h) ‘farmers (are alleged to) have their hand out’. We can then represent our best guess about the time relations pictorially in Figure 6.1, where time is represented by the line running from left to right. For simplicity I’m marking the habitual or possibly habitual events as though they happen just once. This representation obviously makes some assumptions and simplifications: for instance, it’s not certain that event (c) actually finishes before the time at which (6.1) is uttered, (a). This is arguably suggested by the use of was spending rather than spend, but is not certain. Also, as noted above, I’m treating (b) and (d) as single rather than habitual events; I’m also treating (a), (b) and (d) as events that take place each



tense and aspect 69

(b)

(d)

(a) (c)

(e) (f) (g)

(h)

Figure 6.1 Time relations among the events described in (6.1). Time runs from left to right: open-ended boxes indicate events with no endpoint suggested

at a single point in time, ignoring the fact that these events each have a duration of their own. However, the crucial point for our current purposes is that this complicated picture can be expressed using just a few common, and sometimes barely noticeable, morphemes in what is still a reasonably simple sentence. The rest of this chapter discusses more systematically the options that English makes available to us for expressing these kinds of time reference via our grammar. 6.2 Tense

In the names of the verb forms we’ve seen so far – past simple, present perfect, and so on – the first part of the name indicates tense and the second part indicates aspect. Nine different combinations of tense and aspect are set out in Table 6.1. This section will discuss tense in more detail, while Section 6.3 focuses on aspect.

Table 6.1  Two-part labels for tense–aspect combinations, with examples simple aspect progressive aspect perfect aspect

Past tense

Present tense

Future tense

past simple saw

present simple see

future simple will see

past progressive present progressive future progressive was/were seeing am/is/are seeing will be seeing past perfect had seen

present perfect have/has seen

future perfect will have seen

70

an introduction to english semantics and pr agmatics

6.2.1 Preliminaries

As discussed in Section 6.1, a tense form can be used to convey various different time points or intervals, depending on the time of utterance and on other features of the sentence. In this respect, tense is deictic, but the process of understanding the temporal meaning of tense is more complicated than for other deictic expressions, as we have already seen. A time adverbial such as yesterday, right now or in ten minutes’ time anchors an event in the past, present or future, respectively, relative to the time of utterance. Even though we refer to past, present and future tense, the mapping between these expressions and the actual time periods they denote is not always as straightforward as this: sometimes we need more elaborate pragmatic reasoning in order to determine the intended meaning of a tense form. As we will see in Section 6.3, the same is true of aspectual marking. The forms that explicitly encode tense and aspect in English are the explicit markers listed below, although we should also be aware that the “unmarked” forms of verbs (see, look, can, and so on) are also associated with specific kinds of tense and aspect: • Auxiliary verbs: will, have, be. • Irregular forms of verbs: saw, seen, thought, blew, blown, is, am, etc. • Inflectional suffixes: past tense, usually written -ed present tense, usually written -s when the subject is singular, not the speaker and not the hearer progressive -ing, for example am singing, was emerging past participle -(e)n or -ed, for example has seen, have helped We’ll now consider some of the time meanings that English speakers express using these forms. ●







6.2.2  Present, past and future

(6.2) shows some examples of the present tense being used with reference to events and states that occur or exist in a period of time that includes the time of utterance. Hence, we could tag now or at this precise moment onto the end of any of these sentences and the result would be coherent. (6.2) a. He goes for goal. (Said by a sports commentator) b. That dog is happy. c. It’s wagging its tail.



tense and aspect 71

Present forms are also used for so-called “timeless truths”, as in (6.3). Someone uttering one of these sentences is not just making a claim about what happens to be the case at the time of utterance. The adverb always could be added to those utterances without changing the intended meaning, whereas adding at this precise moment would change the meaning in a somewhat odd way. (6.3) a. At sea level, water boils at 100ºC. b. Dark clouds have a silver lining. It makes some sense to use the present tense to talk about timeless truths: after all, they are as true at the present moment as they are at any other time. More surprisingly, we can use the present tense with adverbial markers of future time, such as next year and a week on Wednesday, as shown in (6.4). (6.4) a. You arrive in Australia next month in time for the Melbourne Cup. b. I am giving a talk in Newcastle a week on Wednesday. Both of these sentences are unproblematic – it’s easy to imagine them being uttered in connection with a travel itinerary or a diary entry – and the use of the present tense doesn’t detract from the future time ­reference of the sentence as a whole. We can’t generally do this for past time reference, as shown in (6.5). (6.5) a. *You arrive in Australia last month in time for the Melbourne Cup. b. *I am giving a talk in Newcastle two Wednesdays ago. The only way to make sense of the sentences in (6.5) is to interpret them as part of a narration in which the speaker has switched to using the present tense for narrative effect (sometimes called the historic present tense). One possible explanation for the asymmetry between (6.4) and (6.5) is that Germanic languages tend to have two tenses, past and nonpast (Hewson and Bubenik 1997: 209), so we might think of the English present tense as really being a non-past tense that is appropriate to use in reference to present or future events. For the past tense, the examples in (6.6) demonstrate its typical function of communicating about events and states located prior to the time of utterance: in these examples, the times are indicated by the adverbials at 7 o’clock this morning and yesterday. (6.6) a. We ate at 7 o’clock this morning. b. I heard it on the news yesterday.

72

an introduction to english semantics and pr agmatics

However, as shown in (6.7), the past tense can also be used with reference to events that haven’t (yet) happened but could do so. (6.7) If we introduced proportional representation, there would be more coalition governments. We have already seen that we can refer to future events using the present tense. The nearest thing English has to a grammatical marker of future time is will. This is a modal auxiliary verb, and historically had the sense of ‘want’ or ‘wish’, which naturally refers to a future situation. Subsequently it developed a further sense as a general marker of futurity that doesn’t imply the existence of any wish or preference. (6.8a) does not express something that Kiribati wants to happen, nor does (6.8b) express something that the hearer wants. (6.8) a. A small rise in sea-level and Kiribati will disappear under the Pacific. b. You’ll have to leave if you carry on like that. An alternative way to mark futurity is by using going to. This has undergone a similar grammaticalisation process to will: although it still retains a sense in which it can denote physical motion towards a place, it has a distinct sense in which it conveys futurity without suggesting motion, as in (6.9a). In some cases, such as (6.9b), this results in ambiguity as to which sense the speaker intends to express – motion, or futurity without motion. (6.9) a. He’s going to stay at home and look after the kids. b. I am going to work. By contrast, will is less often ambiguous, although it can also be used in at least one other way: for conveying timeless truths, just as the present simple was in (6.3). In (6.10), for instance, it is clear that the speaker is not making a prediction about the future action of a specific diamond or a specific rising tide. If we replaced will cut with cuts and will lift with lifts, this would make very little difference to the speaker’s meaning in each case. (6.10) a. A diamond will cut glass. b. A rising tide will lift all boats. 6.2.3  Tense and adverbials

Past, present and future time (relative to the time of utterance) can be expressed in sentences using various deictic adverbials, as we have



tense and aspect 73

Table 6.2  The compatibility of some deictic adverbials with past, present and future time Past time

Present time

Future time

then

now

then

last year last Bastille Day yesterday

at present nowadays

next year tomorrow in 45 minutes from now

today, this week, this year

already seen. Table 6.2 outlines how some of these adverbials can be used for different times. The adverbials listed in the ‘future time’ column can be used with present simple and present progressive tense forms, as in (6.11a–c), but the resulting sentences clearly describe future events. The adverbials in the ‘past time’ column can also be used with present tense forms, as in (6.11d), but these have to be interpreted as historic present forms which are talking about past events. (6.11) a. Mark Lawson is here in 45 minutes. (Radio announcer describing the next programme) b. She lectures in Milton Keynes tomorrow. c. He’s visiting Scotland next year. d. Last year, he loses his job. Some deictic adverbials are compatible with past, present or future time reference: these include today, this week and this year. These can all make reference to the time of utterance, a time before the utterance or a time after. Consequently, they can freely combine with all tense forms. 6.3 Aspect

Tense provides pointers to the location of events in time, relative to the time of utterance. Once you have “thought yourself into” the appropriate point in time, aspect comes into play. Aspect is about how we encode how an event occurs in time. Do we treat it as though it is compressed into an instant, or mentally stretch it out? Are we interested in the middle stages of ongoing events, or their culminations? The languages of the world provide various different ways of expressing aspect through the grammar. In this section, we look at three forms of aspect: first, habitual occurrences versus single events, and then two kinds of

74

an introduction to english semantics and pr agmatics

aspect that are explicitly marked in the grammar of English, namely progressive and perfect. 6.3.1  Habituality and simple aspect

We talked a little about habitual interpretation when discussing (6.1), repeated below. (6.1) When I told people I was spending time with farmers, they’d say: how can you stand it, they just complain all day and they’ve always got their hand out. A possible interpretation of (6.1) is that the speaker is recounting what happened on a single occasion during which he told people that he was spending time with farmers. However, it is perhaps more natural to think that the speaker is actually generalising over a set of experiences, and that it was in some sense habitual for him to tell people that he was spending time with farmers, and for those people to respond in the way he describes. The actions that those speakers attribute to farmers – complaining, and having a hand out – are also presented as habitual. The difference is that these are qualified by the adverbials all day and always. So, although there is no grammatical marking that specifies habituality in English, it seems that we are able to recognise it, either for contextual reasons or because of the presence of an appropriate adverbial. Table 6.3 presents a series of sentences, in past, present and future tense, which admit a possible habitual interpretation even without any specific adverbials that point in that direction. For instance, the past tense sentence Tim drank decaf can clearly be a description of a single past event: for example, it could be used to tell us what Tim drank after dinner yesterday. However, it can also be interpreted as a statement about Tim’s past coffee-drinking habits. The same applies to the other examples in Table 6.3. We can force a habitual interpretation by adding Table 6.3  A range of sentences which all have habitual as a possible interpretation Past simple

Present simple

Future simple

Maya loved music.

Maya loves music.

Maya will love music.

Tim drank decaf.

Tim drinks decaf.

Tim will drink decaf.

Little Maurice brushed Little Maurice brushes his teeth by himself. his teeth by himself.

Little Maurice will brush his teeth by himself.



tense and aspect 75

a suitable adverbial – in those days, nowadays, in the future – as in (6.12). But in the absence of an adverbial, both possible interpretations are in play. (6.12) a. Maya loved music in those days. b. Tim drinks decaf these days. c. Little Maurice will brush his teeth by himself in the future. There is also a distinction between the first set of sentences and the other two sets. In the drank decaf and brushed his teeth sentences, we can force a single-event reading by using a regular time adverbial like yesterday, today or tomorrow. This is a bit odd with the loved music sentences. This is because those sentences denote states, rather than activities, a distinction discussed by Vendler (1957): if we say that Maya loves music, we are not pinpointing a single instance of her loving music, but rather commenting that she is in the general condition of loving music. Because of the nature of states, sentences like this convey habitual meanings (or something very like habitual meanings) automatically. A further wrinkle is that the single-event interpretation is not so readily available for the present tense form as the past or future form: (6.13a) is a little odd whereas (6.13b–c) are fine. (6.13) a. ?Tim drinks decaf today. b. Tim drank decaf yesterday. c. Tim will drink decaf tomorrow. When we use a present simple form such as drinks, there is a strong preference for the habitual reading (Miller 2002: 148), with perhaps a few specific exceptions such as sports commentary (as in (6.2a)). But this may be because a single event is usually more naturally described with a different form, in this case is drinking, as in (6.14). The kind of aspect exhibited by is drinking – progressive aspect – is the topic of Section 6.3.2. (6.14) Tim is drinking decaf today. To summarise: all of the simple aspect sentences in Table 6.3 allow a habitual interpretation, and this is the only interpretation available for the three state-denoting sentences in the top row. The other six are open to both habitual and single-event interpretations, but the habitual reading is usually preferred for present simple forms. Habitual interpretations can be made obligatory by the use of certain adverbials, but in the absence of these we have to rely on the context of the utterance (and the other material in the sentence) to understand which of the two possible interpretations is intended by the speaker.

76

an introduction to english semantics and pr agmatics

6.3.2  Progressive aspect

Progressive aspect is marked in English by the use of the verb be (in some appropriate form) and the suffix -ing. The use of progressive aspect presents an event as an ongoing activity which occurs for a protracted period of time – it doesn’t focus on the beginning or the end of the event. In narrative, background events are often presented in a clause with progressive aspect, as in (6.15). (6.15) The rain was falling as I left the house. (, retrieved 26 April 2022) Here, the scene is set by a progressive clause describing the (ongoing) weather, and a single event is presented as occurring against that backdrop, namely the speaker leaving the house. If we wanted to focus on the onset of the weather, we could switch the aspectual marking around, as in (6.16). (6.16) As I was leaving the house, the rain started to fall. Progressive aspect can be used to present even a short event as an ongoing activity, thus making it a possible backdrop for other events. For instance, (6.17) presents the departure of a bus as an ongoing activity, which the hearer is presumably being invited to interrupt with the single momentary action of boarding the bus. (6.17) Hurry, the bus is leaving. We can see how progressive aspect disregards the ends of events by considering the entailments of sentences such as (6.18)–(6.20). Here, the (c) sentences are in progressive form. (6.18) a. The firm demolished the building. b. The firm has demolished the building. c. The firm was demolishing the building. (6.19) a. The waiter folded a napkin. b. The waiter has folded a napkin. c. The waiter was folding a napkin. (6.20) a. They drew up a contract. b. They have drawn up a contract. c. They were drawing up a contract. In each of these cases, we see that the (a) and (b) sentences entail that the relevant event was successfully completed, whereas the (c) ­sentence



tense and aspect 77

does not: it is compatible with a situation in which the action was interrupted. The above examples all involve “accomplishments”, in Vendler’s (1957) taxonomy of events: they have a duration in time and they have a well-defined point of completion. By contrast, verbs that encode states tend to be less compatible with progressive aspect, as shown by (6.21a). Where we can use progressive aspect with state verbs, the effect is sometimes what Cruse (2011) calls “provisionality”: the sense that the state is currently in effect but may not be in the future. For instance, (6.21b) seems to suggest that the hearer’s keys can be retrieved, whereas the non-progressive (6.21c) suggests that the speaker’s hopes cannot. (6.21) a. *Who is knowing Danish? b. Your keys are lying at the bottom of the swimming pool. c. My hopes lie shattered. 6.3.3  Perfect aspect

In English, the combination of the auxiliary verb have and the past participle form of a verb expresses what is called perfect aspect. The perfect (from the Latin perfectum, meaning completed) is used to describe occurrences in the aftermath of an event or state: that is to say, in the period (however long) during which the event or state seems to continue to have direct consequences. To put it another way, when we use the perfect, we convey that the occurrences are in the aftermath of some event or state. This is a slightly elusive notion, but we can illustrate it with some examples. Consider (6.22) and (6.23). (6.22) a. I passed my driving test. b. I have passed my driving test. (6.23) a. I passed my school leaving exams. b. I have passed my school leaving exams. Immediately after passing their driving test or their school leaving exams, a speaker might utter any of these sentences. But if the speaker is looking back years after the event, (6.23b) seems much less appropriate than (6.23a) – even though the event may have been very relevant to the life they live now, its direct consequences no longer reverberate. It is perhaps easier to imagine a context in which (6.22b) would still be appropriate years later – perhaps the speaker is reassuring a nervous passenger – because the fact of their having passed their driving test may still have immediate relevance. As Quirk et al. (1985: 193) put it,

78

an introduction to english semantics and pr agmatics

two common features of present perfect meaning are that ‘the relevant time zone leads up to the present’ and ‘the result of the action still obtains at the present time’. Of course, if we are in the aftermath of an event, this entails that the event has been completed. We see this in the (b) examples of (6.18)–(6.20) as well as those of (6.22) and (6.23). In cases such as (6.24), arguably being in the aftermath of an event tells us something more. (6.24) a. The rain started. b. The rain has started. Both versions of (6.24) entail that the event of ‘the rain starting’ has ­happened. But under normal circumstances (6.24b) also conveys that it is still raining at present, whereas (6.24a) is compatible with a situation in which the rain started and stopped again prior to the time of utterance. Having said that, what counts as the aftermath – and consequently, what a sentence like (6.24b) tells us – is potentially dependent on context. If (6.24b) was spoken by a scientist who predicted that climate change would bring rain to some previously arid desert, it could be used to report signs that rain had fallen there, rather than to report that it is presently raining there. In this case, we’re relying on a habitual interpretation: the speaker is reporting that we are in the aftermath of the change from it habitually not raining to it habitually raining. Linguists have noted that present perfect forms tend not to accept past time adverbial modifiers, as illustrated in (6.25a). However, Klein (1992) pointed out that the present perfect unexpectedly accepts members of a small class of past time adverbials, including recently. The contrast between (6.25b) and (6.25c) illustrates that recently behaves as a past time adverb, but (6.25d) shows that it fits with a present perfect. (6.25e) shows that the same is true for a preposition phrase with since. (6.25)

a. b. c. d. e.

*I have arrived yesterday. *They go there recently. They went there recently. They have been there recently. They have been there since last week.

This pattern apparently arises because of how the meanings of recently and since last week make reference to the time of utterance: the periods they describe run up to the present time, which makes them compatible with the present perfect. Past time adverbials which denote times that do not extend up to the present, such as yesterday, do not work so well.



tense and aspect 79

6.3.4  Perfect aspect or tense?

We have discussed how tense locates events in time, whereas aspect describes the time profile of those events. However, perfect aspect does locate events in time, in that it is used only in the aftermath of those events. Hence, from the rain has started, we infer that the event of ‘the rain starting’ happened before the time of utterance, even if only a moment before. This is one reason why the perfect form is sometimes called the “perfect tense”, as it is in Huddleston and Pullum’s (2002) authoritative grammar of English. In the examples we have seen of the present perfect, we do also have a potential place to mark tense, the verb have. If we replaced have with had, the clauses would be past perfect, as in the underlined parts of (6.26). (6.26) a. When I started work, I had passed my driving test. b. As I left the house, the rain had started. (6.26a) clearly places the event of the speaker starting work in the aftermath of them passing their driving test, and (6.26b) places the event of them leaving the house in the aftermath of the rain starting. In both cases, the first clause deictically points to a location in time before the time of utterance, and the past perfect places its event at a time before that. Summary

Tense is deictic. It locates events in relation to the time of utterance: past, present or future. In English, past forms usually appear with a suffix, present forms unmarked or with an -s, and future forms with various kinds of marking. Time adverbials also help to reveal the mapping between tense forms and time, which can be surprisingly flexible and context-dependent. Aspect is about the time profile of events. The grammatically marked forms in English are progressive (ongoing without attention to the ending) and perfect (we are in, or talking about a time in, the aftermath of the event). Habitual aspect is not grammatically marked in English, but is readily available as an interpretation for numerous classes of sentence. Exercises

1. Table 6.2 presents various kinds of deictic adverbial showing the ­different times – relative to utterance time – that they are ­compatible

80

an introduction to english semantics and pr agmatics

with. Which group does recently belong in? And where does soon belong? You will need to make up sentences and scenarios for past, present and future tense and try them for compatibility with recently and soon. 2. With reference to aspect, discuss the difference in meaning between Arthur’s a tyrant and Arthur’s being a tyrant. 3. A tobacco company told the Czech government that they had saved many millions of dollars because people were dying early. Think of the sentence in italics as part of a newspaper report (and note that the pronoun they refers to the Czech government). Identify the combinations of tense and aspect used in the sentence. If we wanted to draw a diagram like Figure 6.1 to represent the relative timing of the events mentioned in the sentence, what might be ambiguous about the sentence as it stands? 4. Sentences (a) and (b) illustrate “be to verb” as a rather formal way of marking the future. (a) could be used to notify students about a deadline; (b) could be used on the deadline day to remind them. (a) On 11 May you are to submit a written solution to the exercise. (b) You were to submit written solutions today. Sentence (b) embeds a future tense within the past; were is a past tense form and “be to verb” is, as illustrated in (a), a way of marking future.   Now try to find some less formal ways of embedding a future in the past. Suppose your friend promised you (yesterday) that they would lend you a book that you asked for, but they have now forgotten which book. To remind them, you could use a ‘future in the past’ form: past because the promise was made yesterday, future because the book was to be lent at a then-future date. Suggest two reasonable completions for (c) that involve a form of future marking with past tense on a verb. (c) You said you _______________ lend me One Day. How do these compare with the words that you might have used in asking to borrow the book in the first place? Recommendations for reading

Trask (1993) is a good first resort for looking up terms such as tense, aspect, progressive and perfect that may be unfamiliar. Chapter 13 of Miller



tense and aspect 81

(2002) is a short, clear introduction to the meanings associated with tense and aspect. Kearns’s (2011) account of English tense and aspect is also highly accessible and systematic, and there are good discussions in Cruse (2011) and Saeed (2015). Worthwhile generalisations, as well as many interesting details, are available via the index entries for tense and aspect in Huddleston and Pullum (2002) and Quirk et al. (1985).

7 Modality, scope and quantification

Overview

So far we have mostly discussed sentences that categorically state facts about the world: that is to say, they assert that something is true. In practice, we often want to communicate less categorically than this: for instance, to assert that something is possibly true, or that it ought to be true, or that it is permissible for it to be true. We use the term modality to refer to the ways in which we can use language to qualify statements of fact like this. English provides a number of ways of expressing modality, including through the use of modal verbs and adverbial expressions. Modality is important to us in communication, but the expression of modality often gives rise to ambiguity. One way in which this can happen is when a modal expression interacts with another expression, such as negation – in this case, we need to understand the scope relations between the two expressions, which we will discuss in Section 7.2. Finally, we will look at how we can use language to quantify over things. As we shall see, quantification has a lot in common with modality, not least that quantifiers can also enter into complicated and sometimes ambiguous scope relations. 7.1 Modality

To put it very generally, a clause characterises a situation, and modality expresses how the speaker relates to that situation. The underlined expressions in (7.1) contribute modal meanings. (7.1) a. Bill has to apologise. b. You can come in now. c. She’s not able to see you until Tuesday. d. Working so many hours a week, he must be exhausted. 82





modalit y, scope and quantification 83

e. If the pub is open, there ought to be someone behind the bar. f. Martians could be green.

In (7.1a), the verb has to qualifies the situation ‘Bill apologises’, and expresses that there is an obligation to make that situation come about. Can in (7.1b) similarly indicates that a situation (‘the hearer comes in’) is permissible, and able to in (7.1c) indicates that it is feasible. In (7.1d–f), the verbs signal how confident the speaker is in their knowledge about the relevant proposition: must indicates that, in the light of available information, the speaker considers the proposition (in (7.1d), ‘he is exhausted’) to be certainly true; ought to indicates that they consider it likely to be true; could indicates that they consider it possibly true. The main carriers of modality are a set of auxiliary verbs called modals: will, would, can, could, may, might, shall, should and must. However, modality is also encoded by various other expressions, such as possibly, probably, have (got) to, need to, ought to and be able to. 7.1.1  Modal verbs and tense

In Chapter 6, we saw that tense forms are not an entirely reliable indicator of time: it is possible to use present forms when narrating past events, and past forms for future events in conditional sentences, and so on. For modal verbs, past forms are often used in connection with future events. As shown by (7.2a–b), would behaves as the past form of will: but as shown by (7.2c–d), either can be used to make a request that refers to the future. (7.2)

a. b. c. d.

*In the past, John will help me. In the past, John would help me. Will you help me tomorrow? Would you help me tomorrow?

Can and could pattern the same way. When used to make requests about the future, there is very little difference in meaning, although we might intuit that the past tense forms are somewhat more tentative and perhaps more polite than their present tense counterparts. In Chapter 6 we discussed how will can signal future time reference but can also be used to state timeless truths like (7.3a). We can compare this with (7.3b), in which we replace will with a weaker modal, can. (7.3) a. A diamond will cut glass. b. A diamond can cut glass. If we think of will in (7.3a) as a modal, we can understand the sentence as expressing a high level of confidence on the part of the speaker that,

84

an introduction to english semantics and pr agmatics

if you attempt to cut glass with a diamond, you will succeed. (7.3b) expresses a slightly weaker meaning, along the lines that if you do this, you may possibly succeed. But there doesn’t seem to be a difference between (7.3a) and (7.3b) in terms of the time reference they express: will does not contribute a future tense meaning to (7.3a). Thus, when we interpret will, we have to consider whether it is contributing tense or modality. In fact, marking of modality tends more generally to be complementary to tense marking: English syntax often forces us to choose whether a clause will have tense in it or modality instead. The default pattern for English is for clauses to have tense marking but no marking for modality. Expressions of modality exhibit a rich array of useful meanings. As a consequence, the modal auxiliaries are among the most frequently used verbs in English: six of the top twenty verbs in English are will, would, can, could, may and should, each averaging more than 1,000 occurrences per million words of running text (Leech et al. 2001: 282). Biber et al. (1999: 456) analysed a corpus of text samples totalling 40 million words, from a range of genres, and found that modals were used in about 15 per cent of the clauses that could have them. In short, modality is an important topic in English semantics and pragmatics. Whole books have been written about English modals and modality, for example Palmer (1990). Against this backdrop, the aim of this section is naturally quite modest: we will just look at a couple of the principal issues that make modality an interesting topic to study, along with the terminology that is used to discuss these issues. 7.1.2  Epistemic and deontic modality

Recall that examples (7.1a–c) involved modality modifying how situations were presented, whereas (7.1d–f) used modality to qualify the strength of the speaker’s belief in the proposition being discussed. This distinction broadly corresponds to the distinction between what is called – borrowing terms from philosophy – deontic and epistemic modality. Epistemic meaning has to do with knowledge and understanding. Markers of epistemic modality are understood to be expressing the strength of someone’s certainty about the truth of a proposition. Examples (7.4)–(7.7) present pairs of sentences (a) without and (b) with modal markers, which are underlined. (7.4) a. The whole hillside is slipping down into the valley. b. The whole hillside could be slipping down into the valley.



modalit y, scope and quantification 85

(7.5) a. They meet in the final tomorrow. b. They may meet in the final tomorrow. (7.6) a. Jessica went by motorbike. b. Jessica probably went by motorbike. (7.7) a. The van was heavily laden, judging by the tyre marks. b. The van must have been heavily laden, judging by the tyre marks. Speakers can produce modally unqualified sentences, like the (a) examples here, when they are confident in their facts; however, the possibility of marking epistemic modality makes it easy for a speaker to make a less definitive claim if they are not certain, as in the (b) examples. Epistemic modality comes in different strengths: (7.4b)–(7.7b) represent a gradient from weak to strong modality. In (7.4), replacing is by could be considerably weakens the claim that the speaker is making: in (7.4b) they are merely saying that the whole hillside slipping into the valley is not ruled out by the available evidence. At the other end of the scale, in (7.7b), replacing was by must have been makes only quite a subtle difference to the speaker’s commitment to the claim that the van was heavily laden. Deontic meaning is about concepts such as duty, morality, laws and rules. Deontic modality allows speakers to express whether a proposition relates to a situation that is obligatory, permissible, or somewhere in between, as shown by the examples in (7.8), with modal verbs underlined. (7.8) a. You can borrow my bike any time. b. The consul could have been more helpful. c. You should email Tom. d. Tax returns must be submitted by the end of September. (7.8a) shows a common way of giving permission (using may instead of can would also work, and is sometimes preferred prescriptively): the speaker conveys that the situation in which the hearer borrows their bike at any time is a permissible one, as far as they are concerned. In (7.8b), the speaker uses could to convey the idea that it would have been permissible for the consul to be more helpful (and thereby suggests that they would have preferred the consul to be more helpful). The use of should in (7.10c) conveys that the speaker takes the situation in which the hearer emails Tom to be a preferable one. And must in (7.10d) conveys that the timely submission of a tax return is an obligation.

86

an introduction to english semantics and pr agmatics

As the same verbs can be used to express both epistemic and deontic modality, it is quite possible for a sentence to be ambiguous between these interpretations. In fact, some of the examples we have already discussed are ambiguous in principle. An example like (7.9) makes this more obvious. (7.9) After the exam, Amy may take a cigarette break. If (7.9) is uttered by someone who has authority over Amy’s smoking, it is natural to interpret it as a grant of permission: that is, to take may to express deontic modality. If it is uttered by someone who does not have such authority, it is natural to interpret it as a prediction that it is possible that Amy will take a cigarette break after the exam – and in this case we are taking may to express epistemic modality. In interpreting (7.9), we rely on context to establish whether the modal expression is meant to be epistemic or deontic: specifically, we rely on facts about the speaker and their likely authority. And often it is reasonably straightforward to disambiguate modal expressions in this way. Consider (7.10)–(7.12), in which the modal expressions are underlined. (7.10) a. It may be dark by the time we’ve finished. b. If you wish, you may copy these two diagrams. (7.11) a. Random numbers can appear to have patterns in them. b. The pigeons can have this bread. (7.12) a. At 95 metres, this has got to be one of the tallest trees in the world. b. He has got to be more careful or he’ll break something. In the (a) examples, the speaker is discussing natural phenomena, and it makes no sense for the speaker to attempt to grant permission for night to fall, or for random numbers to exhibit apparent patterns. Nor does it make any sense to place an obligation on a tree to be one of the tallest in the world. These potential deontic meanings are sufficiently absurd that we can dismiss them even without consciously entertaining them: we naturally seem to understand the (a) sentences as expressing epistemic modality. By contrast, the (b) sentences do seem to express authority on the part of the speaker, and are naturally compatible with deontic interpretations. The ambiguity between epistemic and deontic modality also arises in cases that do not involve core modal verbs. (7.13) provides an example of this: the underlined content expresses epistemic modality in the context of (a) and deontic modality in the context of (b). We are led to



modalit y, scope and quantification 87

this understanding by the surrounding context: in (a), Marie’s expectation is based on her prior experience, whereas in (b) it is an expression of how she thinks things ought to be. (7.13)

a. Marie expected the coffee to be strong; she had ordered it before. b. Marie told the waiter that she expected the coffee to be strong and that she would send it back if it wasn’t.

Under certain circumstances, the ambiguity between deontic and epistemic modality may be resolved incorrectly, and this can have important consequences. The physicist Richard Feynman identified an example of this while working on the inquiry into the failure of the Space Shuttle Challenger in 1986. He quoted NASA officials as making the argument presented here as (7.14). (7.14) Since the Shuttle is a manned vehicle, the probability of mission success is necessarily very close to 1.0. (Feynman 1999: 154) The underlined material in (7.14) admits two possible interpretations. On an epistemic interpretation, it affirms that, given the available evidence, the mission could be judged to be almost certain to succeed. On a deontic interpretation, it affirms that the mission must be almost certain to succeed: that is, if it isn’t almost certain to succeed, it should not take place. Feynman noted that NASA’s statement was intended to be deontic, but was widely misinterpreted as epistemic, giving rise to an irrational belief that the Shuttle mission was a lot safer than it actually was. 7.2  Semantic scope

In the preceding section, we saw how ambiguity could arise between two possible readings of modal expressions, one epistemic and one deontic. But additional ambiguities arise with the use of modals, particularly when combined with certain other kinds of expression, such as negation. In Section 2.2, we discussed how words such as unlockable can systematically have two meanings, one in which the prefix un- combines with the word lockable to give the meaning ‘not able to be locked’, and one in which the suffix -able combines with the word unlock to give the meaning ‘able to be unlocked’. We could describe these different derivational processes by using brackets, and distinguishing (un(lockable)) from ((unlock)able). Another way of looking at this is to appeal to the idea of scope: that is to say, the material that a particular operator applies to. We could think

88

an introduction to english semantics and pr agmatics

of the prefix un- and the suffix -able as operators on the meanings of other lexical items, which achieve the effects respectively of ‘reversing the direction of change in a verb meaning’ and ‘turning a verb into an adjective of possibility’. Then we could say that the difference between (un(lockable)) and ((unlock)able) is that, in the former, un- takes scope over -­ able, while in the latter the reverse is true. So in the former case the resulting word meaning denies the possibility of something being locked, whereas in the latter case the resulting word meaning expresses the possibility of something being unlocked. The general point here is that, when we have two operators in the same expression, we can get different meanings depending on which operator includes the other within its scope. This is an important issue for modality. Consider the deontic interpretations of the sentences in (7.15). (7.15)

a. b. c. d.

You must discuss the case. You have to discuss the case. You mustn’t discuss the case. You don’t have to discuss the case.

(7.15a) and (7.15b) are virtually identical in meaning, but their negated forms (7.15c) and (7.15d) are sharply different: (7.15c) conveys that discussing the case is forbidden, whereas (7.15d) merely conveys that it is not obligatory. We can think of these meanings as involving two operators: negation, expressed here by n’t, and obligation, expressed here by must and have to. In (7.15c), obligation takes scope over negation, and the sentence consequently expresses an obligation towards a negated proposition (‘hearer does not discuss the case’). In (7.15d), negation takes scope over obligation, and the sentence expresses a lack of obligation towards a positive proposition (‘hearer discusses the case’). We can write these scope relations down using brackets, as shown in (7.24). (7.16) a. obligatory (not (you discuss the case)) b. not (obligatory (you discuss the case)) In this example we have two very similar modal expressions conveying different meanings because they interact differently with negation. In (7.17), we can see two very different modal expressions conveying very similar meanings – again, because they interact differently with negation. (7.17) a. They must not have received the invitation. b. They can’t have received the invitation.



modalit y, scope and quantification 89

Either of these sentences could be used to convey an epistemic modal meaning, which might be intended to explain why some expected guests have failed to show up at an event. But the modals must and can obviously differ in meaning. It is their interaction with the negation that cancels out this difference. In (7.17a), just as in (7.15c), must takes scope over the negation, with the resulting meaning that ‘it is certain that they did not receive the invitation’. But in (7.17b), can falls within the scope of the negation (as it typically does), with the resulting meaning ‘it is not possible that they received the invitation’. And this is essentially the same meaning as that yielded by (7.17a). Negation in English is complicated to interpret because its semantic scope is not entirely predictable from syntax alone. As a first approximation, we might guess that negation will take scope over a modal operator if the negation comes first in the sentence: but (7.17b) shows us that this is not always true. A more refined idea would be that the scope relations of operators have something to do with their relative positions in the syntactic tree: however, as we shall see in the following section, semantic ambiguities seem to arise from some sentences that appear to have unambiguous syntactic analyses. In addition to this, English has a widespread class of expressions, sometimes referred to as “neg-raising predicates” (Horn 1989), which cause negation to be interpreted in a way that doesn’t straightforwardly match the syntax. (7.18) presents some examples of this. (7.18)

a. b. c. d.

I don’t think Bill is going to win. Elena doesn’t seem happy. Tom doesn’t want to move to London. Aliya doesn’t intend to take that risk.

The speaker of (7.18a) appears to be saying that ‘it is not the case that I think Bill is going to win’: that is, they are expressing the negation of ‘I think Bill is going to win’. However, in practice, we understand them to be conveying ‘I think Bill is not going to win’: that is, we actually interpret the negation as taking scope over going to win rather than think Bill is going to win. Similarly, (7.18b–d) are all naturally interpreted as making stronger statements than they appear to: we understand the speakers to be saying that ‘Elena seems unhappy’, ‘Tom wants not to move to London’ and ‘Aliya intends not to take that risk’. In general, then, we can’t just read scope relations off the syntax: we are going to have to combine syntactic information, individual word meanings, and contextual information in order to understand what the speaker meant to convey. This is true when we combine modality with other operators, when we interpret negation, and when we combine

90

an introduction to english semantics and pr agmatics

various other kinds of operator, including those we will discuss in the next section. 7.3 Quantification

English provides a range of expressions that we can use to provide information about the number, or proportion, of individuals that have certain specific properties. These expressions include quantifiers such as some, all, none, many and most, as well as numbers, modified numerals such as more than 10, and so on. We can think of quantified sentences such as those in (7.19) as making statements about the number of distinct entities in sets of things. In (7.19a), this is the set of cats that lack tails; in (7.19b), it is the set of men who have walked on the Moon; in (7.19c), it is the set of countries in Africa. (7.19) a. Some cats lack tails. b. Twelve men have walked on the Moon. c. There are more than fifty countries in Africa. For this reason, before we discuss the meanings of quantifying expressions in general, it will be useful to introduce some ideas about sets as used in semantics and pragmatics. 7.3.1  Some basics about sets

A set is a collection of distinct objects, which can be of any kind. The objects in the set are referred to as the elements or members of the set. We can define a set either by listing all of its members or specifying the properties that its members must have in order to qualify as part of the set. (This latter option will be more relevant to our needs here.) A given element can be a member of multiple sets at the same time. It’s customary to denote sets by capital letters. If two sets A and B have exactly the same members, we can say that the sets are identical: we can write this as A = B. If every member that is an element of one set, C, is also an element of some other set, D, then we can say that C is a subset of D, and write this as C ⊆ D. For instance, if C is the set of numbers that are divisible by 9 and D is the set of numbers that are divisible by 3, every member of C is automatically a member of D, but the reverse is not true. We write the empty set – the set with no members – with the symbol ∅. (Technically we say that C is a “proper subset” of D if C ⊆ D, C is not equal to D, and C is not equal to ∅.) We use the word size or cardinality to describe the number of dis-



modalit y, scope and quantification 91

tinct, non-identical members in a set. The empty set ∅ has cardinality zero. If A = B, both A and B have the same cardinality. We can write these relations as |∅| = 0 and |A| = |B|. We can combine sets in particular ways to form other sets. For instance, we can define the union of two sets, E and F, as the set comprising all the entities that are members of E or members of F (or both). We write this as E ∪ F. We can define the intersection of two sets, E and F, as the set comprising all the entities that are both members of E and members of F. We write this as E ∩ F. As these definitions suggest, union and intersection are both symmetric operations: E ∪ F = F ∪ E, and E ∩ F = F ∩ E. Because of how these sets are defined, we can write down some statements about the cardinalities of unions and intersections relative to the cardinalities of the original sets: • |E ∪ F| ≤ |E| + |F|. The biggest that the union of two sets can be is the cardinalities of those two sets added together. The union of the sets will only be this big if the sets don’t overlap at all – that is, there is no single entity that is a member of both sets – in which case we say that the sets are disjoint. In that case, the intersection will be empty: E ∩ F = ∅. • |E ∩ F| < |E|, |E ∩ F| < |F|. The intersection of two sets can’t be bigger than either of the sets individually (because it can only contain entities that are members of both sets). We can also define the complement of a set as all the things that are not in the set. For the set G, we’ll write the complement as G′. For any two sets G and H, the following relations will be true: • G = (G ∩ H) ∪ (G ∩ H′) • H = (H ∩ G) ∪ (H ∩ G′) That is to say, given any set H, we can divide any set G completely into two parts: the members of G that are also members of H, and the members of G that are not also members of H. However, one of these parts may be empty: if G is a subset of H, then G ∩ H′ is empty. If G and H are disjoint, then G ∩ H is empty. In the following section, we’ll see how we can use some of this theoretical machinery to make sense of how we express quantity in English. 7.3.2  Simple quantifiers in terms of sets

Suppose we want to convey information about whether members of one set have a particular property: for instance, whether pandas are

92

an introduction to english semantics and pr agmatics

P

V Meaning of (7.20a)

V

P

Meaning of (7.20c)

Figure 7.1 Venn diagrams for the meanings of (7.20a) and (7.20c)

vegetarian. Depending on what we want to say, we might use one of the sentences in (7.20). (7.20) a. No pandas are vegetarian. b. Some pandas are vegetarian. c. All pandas are vegetarian. Let P denote the set of pandas and V denote the set of individuals that are vegetarian. (Here I’m using vegetarian in the sense of ‘being an animal that does not eat meat’ rather than the more specific sense of ‘being a person who does not eat meat’.) We can think of the meanings of the sentences in (7.20) in terms of what they say about the relationship between the sets P and V. (7.20a) means that there are no members of P that are also members of V: to put it more formally, that the intersection of P and V is empty, which we can write as P ∩ V = ∅. (7.20c) means that every member of P is also a member of V: that is, that P is a subset of V, which we can write as P ⊆ V. Figure 7.1 illustrates these two meanings with Venn diagrams. (7.20b) is a little bit trickier. Minimally, we could say that this sentence means that there exists a member of P that is also a member of V: that is, that the intersection of P and V is non-empty. However, if there existed only one vegetarian panda in the world, you might judge (7.20b) to be false on the grounds that ‘one’ is hardly ‘some’. So we might instead say that (7.20b) means that there are multiple members of P that are also members of V: the intersection of P and V has cardinality at least two (or at least three). For now I’ll leave that controversy aside and assume that some just requires the intersection of the sets to be non-empty (that



modalit y, scope and quantification 93

is, ‘there exists a vegetarian panda’). We’ll also ignore the fact that there is a sense of some in which it is just a plural indefinite determiner, as in I ate some cakes: that is not the sense that is relevant for (7.20b). If we assume that the above analysis is correct, then we can write down the meanings of (7.20) in terms of the set relations in (7.21). (7.21) a. P ∩ V = ∅ b. |P ∩ V| ≥ 1 (or P ∩ V ≠ ∅) c. P ⊆ V This formalism makes it relatively easy to establish what the entailments for these sentences are. For instance, we can see that – assuming that pandas exist at all – (7.21c) entails (7.21b), because if P ⊆ V then |P ∩ V| ≥ 1, unless P = ∅. So, according to this formalism, if all pandas are vegetarian is true, then some pandas are vegetarian is also true. In practice, people are split on how they judge the truth of sentences with some in situations where the corresponding sentence with all is true. There is an intuition that some actually means ‘some but not all’ and the above semantic analysis doesn’t capture that. However, there are good reasons to suppose that the semantic meaning of some is just that shown above, and that the meaning ‘some but not all’ arises because of a pragmatic enrichment. We’ll see a possible explanation of why this happens in Section 8.2. We can provide a similar analysis for numerical quantifiers, as in the examples in (7.22). Rather than talking about pandas that are vegetarian – it seems odd to try to count the number of pandas in the world that are vegetarian – we’ll discuss some examples concerning how many pandas are in the zoo. (7.22) a. There are three pandas in the zoo. b. There are more than three pandas in the zoo. c. There are at most ten pandas in the zoo. Let P denote the set of pandas and Z denote the set of animals in the zoo. A straightforward semantic analysis of the sentences in (7.22) is proposed in (7.23). (7.23) a. |P ∩ Z| = 3 b. |P ∩ Z| > 3 c. |P ∩ Z| ≤ 10 On this analysis, (7.22a) means that there are exactly three distinct ­entities that are members of the set of pandas and that are also members of the set of zoo animals. (7.22b) means that there are more than three distinct entities that belong to both these sets, and (7.22c)

94

an introduction to english semantics and pr agmatics

means that there are ten or fewer distinct entities that belong to both these sets. Unfortunately, in practice, the meanings of numerical quantifiers in English are somewhat more complicated than this. Even plain, unmodified numerals, like three in (7.22a), are ambiguous: the sentence might mean that the number of pandas in the zoo is exactly equal to three (the analysis given in (7.23a)), or it might just mean that there exist three pandas that are in the zoo, which might be the intended interpretation if the speaker has something to say about these three pandas in particular. In the latter case, the total number of pandas in the zoo might actually be more than three, in which case the analysis given in (7.23a) doesn’t apply. And it’s a similar story for expressions such as more than, at least, up to, and so on. We can easily write down formalisms that capture their mathematical meaning – we use > for more than in (7.23b), ≤ for at most in (7.23c), and so on – but actually the way they are used in natural language has some additional subtleties to it. However, for our present purposes, we can assume that the analyses presented above are reasonably good approximations to the meanings of these English-language expressions. 7.3.3  Proportional quantifiers

Most of the quantifiers discussed in the preceding section have a symmetry to their meaning. To say that Some pandas are vegetarian is equivalent, in terms of set membership, to saying that Some vegetarians are pandas. Similarly, saying that There are more than three pandas in the zoo is equivalent to saying that More than three of the things in the zoo are pandas. This is simply because, for any sets A and B, A ∩ B = B ∩ A, which follows from the definition. However, not all quantifiers possess this kind of symmetry. The notable exception in Section 7.3.2 is all: all pandas are vegetarian does not mean the same thing as all vegetarians are pandas. In terms of set membership, this is because the relationships C ⊆ D and D ⊆ C are not equivalent – in fact, the only time they can both be true at once is if C = D. So the only sentences with all which can be inverted in this way are the ones in which the two sets being referred to have precisely the same members, as in (7.24). (7.24) a. All even numbers are divisible by two. b. All numbers divisible by two are even. We could think of all as a special case of a larger class of items called proportional quantifiers, which express the proportions of sets that have particular properties. Some other examples are given in (7.25).



modalit y, scope and quantification 95

(7.25) a. Most pandas are vegetarian. b. Less than 50 per cent of pandas are vegetarian. c. Few pandas are vegetarian. In terms of sets, we can think of these as making statements about the proportion of members of the set P that are also members of the set V. (7.25b) is perhaps the most clear-cut example: it appears to mean that less than 50 per cent of the members of P are also members of V. We can formalise that meaning in terms of sets in various ways. For instance, recall that everything in P is either a member of V or a member of its complement, V′, the set of animals that are not vegetarians. If less than 50 per cent of the members of P are members of V, more than 50 per cent of the members of P must therefore be members of V′. This is equivalent to saying that there are more things in the intersection of P and V′ than there are in the intersection of P and V. So we can express the meaning of (7.25b) as (7.26) – put into different words, ‘the number of non-vegetarian pandas exceeds the number of vegetarian pandas’. (7.26) |P ∩ V′| > |P ∩ V| We could say something very similar for (7.25a): most seems to mean ‘more than half’, in which case we could write this meaning down as (7.27): ‘the number of vegetarian pandas exceeds the number of non-vegetarian pandas’. In practice, there are reasons to think that the meaning of most is a bit more complicated than that: for instance, it seems odd to say that Most Americans are female, but that statement is factually correct if most simply means ‘more than 50 per cent’ (see Solt 2011). (7.27) |P ∩ V| > |P ∩ V′| Example (7.25c) is also tricky to define precisely. Few in this sense does not appear to refer to a stable number, or range of numbers: it appears to be context-dependent. For instance, it appears to denote a much smaller number in (7.28a) than in (7.28b). (7.28) a. Few people have been into space. b.  Few people voted in the elections to the European Parliament. One way to approach this might be to propose that few requires the set of individuals with the property in question to be much smaller than the set of individuals without that property: that is, in (7.25c), that vegetarian pandas are greatly outnumbered by non-vegetarian pandas. We could write this down with the expression in (7.29), where