Afroasiatic languages

Source: Wikipedia, the free encyclopedia.
(Redirected from
Afro-Asiatic languages
)

Afroasiatic
Hamito-Semitic, Semito-Hamitic, Afrasian
Geographic
distribution
North Africa, West Asia, Horn of Africa, Sahel, and Malta
Linguistic classificationOne of the world's primary language families
Proto-languageProto-Afroasiatic
Subdivisions
ISO 639-2 / 5afa
Glottologafro1255
Distribution of the Afroasiatic languages

The Afroasiatic languages (or Afro-Asiatic, sometimes Afrasian), also known as Hamito-Semitic or Semito-Hamitic, are a language family (or "phylum") of about 400 languages spoken predominantly in West Asia, North Africa, the Horn of Africa, and parts of the Sahara and Sahel.[2] Over 500 million people are native speakers of an Afroasiatic language, constituting the fourth-largest language family after Indo-European, Sino-Tibetan, and Niger–Congo.[3] Most linguists divide the family into six branches: Berber, Chadic, Cushitic, Egyptian, Semitic, and Omotic.[4] The vast majority of Afroasiatic languages are considered indigenous to the African continent, including all those not belonging to the Semitic branch.

Sidaama language, the Semitic Tigrinya language and the Omotic Wolaitta language, though most languages within the family are much smaller in size.[5] There are many well-attested Afroasiatic languages from antiquity that have since died or gone extinct, including Egyptian and the Semitic languages Akkadian, Biblical Hebrew, Phoenician, Amorite, and Ugaritic. There is no consensus among historical linguists as to precisely where or when the common ancestor of all Afroasiatic languages, known as Proto-Afroasiatic, was originally spoken. However, most agree that the Afroasiatic homeland was located somewhere in northeastern Africa, with specific proposals including the Horn of Africa, Egypt, the eastern Sahara. A significant minority of scholars argues for an origin in the Levant. The reconstructed timelines of when Proto-Afroasiatic was spoken vary extensively, with dates ranging from 18,000 BC to 8,000 BC. Even the latest plausible dating makes Afroasiatic the oldest language family accepted by contemporary linguists.[6]

Comparative study of Afroasiatic is hindered by the massive disparities in textual attestation between its branches: while the Semitic and Egyptian branches are attested in writing as early as the

middle (t-), causative (s-), and passive
(m-) verb forms (Semitic, Berber, Egyptian, Cushitic), and a suffix used to derive adjectives (Egyptian, Semitic).

Name

In current scholarship, the most common names for the family are Afroasiatic (or Afro-Asiatic), Hamito-Semitic, and Semito-Hamitic.[9][10] Other proposed names that have yet to find widespread acceptance include Erythraic/Erythraean, Lisramic, Noahitic, and Lamekhite.[11][12]

Elamites are ascribed to Shem despite their language being totally unrelated to Hebrew.[15] The term Semitic for the Semitic languages had already been coined in 1781 by August Ludwig von Schlözer, following an earlier suggestion by Gottfried Wilhelm Leibniz in 1710.[16] Hamitic was first used by Ernest Renan in 1855 to refer to languages that appeared similar to the Semitic languages, but were not themselves provably a part of the family.[16] The belief in a connection between Africans and the Biblical Ham, which had existed at least as far back as Isidore of Seville in the 6th century AD, led scholars in the early 19th century to speak vaguely of "Hamian" or "Hamitish" languages.[17]

The term Hamito-Semitic has largely fallen out of favor among linguists writing in English, but is still frequently used in the scholarship of various other languages, such as German.

monophyletic "Hamitic" branch exists alongside Semitic. In addition, Joseph Greenberg has argued that Hamitic possesses racial connotations, and that "Hamito-Semitic" overstates the centrality of the Semitic languages within the family.[20][21][22] By contrast, Victor Porkhomovsky suggests that the label is simply an inherited convention, and doesn't imply a duality of Semitic and "Hamitic" any more than Indo-European implies a duality of Indic and "European".[12] Because of its use by several important scholars and in the titles of significant works of scholarship, the total replacement of Hamito-Semitic is difficult.[22]

While Greenberg ultimately popularized the name "Afroasiatic" in 1960, it appears to have been coined originally by

Igor Diakonoff in 1980. At present it predominantly sees use among Russian scholars.[21][12]

The names Lisramic—based on the Afroasiastic root *lis- ("tongue") and the Egyptian word rmṯ ("person")—and Erythraean—referring to the core area around which the languages are spoken, the Red Sea—have also been proposed.[9]

Distribution and branches

A diagram of the six widely recognized branches of the Afroasiatic family, including some of the larger or more culturally significant languages in each branch.

Scholars generally consider Afroasiatic to have between five and eight branches. The five that are universally agreed upon are Berber (also called "Libyco-Berber"), Chadic, Cushitic, Egyptian, and Semitic.[23] Most specialists consider the Omotic languages to constitute a sixth branch.[4] Due to the presumed distance of relationship between the various branches, many scholars prefer to refer to Afroasiatic as a "linguistic phylum" rather than a "language family".[24]

M. Victoria Almansa-Villatoro and Silvia Štubňová Nigrelli write that there are about 400 languages in Afroasiatic;[2] Ethnologue lists 375 languages.[3] Many scholars estimate fewer languages; exact numbers vary depending on the definitions of "language" and "dialect".[24]

Berber

The Berber (or Libyco-Berber) languages are spoken today by perhaps 16 million people.[25] They are often considered to constitute a single language with multiple dialects.[26] Other scholars, however, argue that they are a group of around twelve languages, about as different from each other as the Romance or Germanic languages.[27] In the past, Berber languages were spoken throughout North Africa except in Egypt;[28] since the 7th century CE, however, they have been heavily affected by Arabic and have been replaced by it in many places.[29][30]

There are two extinct languages potentially related to modern Berber.[31] The first is the Numidian language, represented by over a thousand short inscriptions in the Libyco-Berber alphabet, found throughout North Africa and dating from the 2nd century BCE onward.[30] The second is the Guanche language, which was formerly spoken on the Canary Islands and went extinct in the 17th century CE.[31] The first longer written examples of modern Berber varieties only date from the 16th or 17th centuries CE.[32]

Chadic

Chadic languages number between 150 and 190, making Chadic the largest family in Afroasiatic.

Chad basin, with the exception of Hausa.[35] Hausa is the largest Chadic language by native speakers, and is spoken by a large number of people as a lingua franca in Northern Nigeria.[36] It may have as many as 80 to 100 million first and second language speakers.[33] Eight other Chadic languages have around 100,000 speakers; other Chadic languages often have few speakers and may be endangered of going extinct.[37] Only about 40 Chadic languages have been fully described by linguists.[33]

Cushitic

There are about 30 Cushitic languages,

Sidaama.[38] Many Cushitic languages have relatively few speakers.[39] Cushitic does not appear to be related to the written ancient languages known from its area, Meroitic or Old Nubian.[41] The oldest text in a Cushitic language probably dates from around 1770;[38] written orthographies were only developed for a select number of Cushitic languages in early 20th century.[39]

Egyptian

Seal impression from the tomb of Seth-Peribsen (c. 2690 BCE), containing the first complete sentence in Ancient Egyptian.[42]

The Egyptian branch consists of a single language, Ancient Egyptian, which was historically spoken in the lower Nile Valley.

hieroglyphs, which only represent consonants.[44] In the Coptic period, there is evidence for six major dialects, which presumably existed previously but are obscured by pre-Coptic writing; additionally, Middle Egyptian appears to be based on a different dialect than Old Egyptian, which in turn shows dialectal similarities to Late Egyptian.[45] Egyptian was replaced by Arabic as the spoken language of Egypt,[46] but Coptic continues to be the liturgical language of the Coptic Orthodox Church.[47]

Omotic

The c. 30 Omotic languages are still mostly undescribed by linguists.[35] They are all spoken in southwest Ethiopia except for the Ganza language, spoken in Sudan.[48] Omotic is typically split into North Omotic (or Aroid) and South Omotic, with the latter more influenced by the Nilotic languages; it is unclear whether the Dizoid group of Omotic languages belongs to the Northern or Southern group.[35] The two Omotic languages with the most speakers are Wolaitta and Gamo-Gofa-Dawro, with about 1.2 million speakers each.[39]

A majority of specialists consider Omotic to constitute a sixth branch of Afroasiatic.[4] Omotic was formerly considered part of the Cushitic branch;[49] some scholars continue to consider it part of Cushitic.[50] Other scholars have questioned whether it is Afroasiatic at all, due its lack of several typical aspects of Afroasiatic morphology.[4]

Semitic

There are between 40 and 80 languages in the Semitic family.[51] Today, Semitic languages are spoken across North Africa, West Asia, and the Horn of Africa, as well as on the island of Malta, making them the sole Afroasiatic branch with members originating outside Africa.[52][51] Arabic, spoken in both Asia and Africa, is by far the most widely spoken Afroasiatic language today,[2] with around 300 million native speakers, while the Ethiopian Amharic has around 25 million.[5]

Most authorities divide Semitic into two branches: East Semitic, which includes the extinct Akkadian language and West Semitic, which includes Arabic, Aramaic, the Canaanite languages (including Hebrew), as well as the Ethiopian Semitic languages such as Ge'ez and Amharic.[53] The classification within West Semitic remains contested. The only group with an African origin is Ethiopian Semitic.[51] The oldest written attestations of Semitic languages come from Mesopotamia, Northern Syria, and Egypt and date as early as c. 3000 BCE.[52]

Other proposed branches

There there are also other proposed branches, but none has so far convinced a majority of scholars:[10]

  • Linguist H. Fleming proposed that the near-extinct Ongota language is a separate branch of Afroasiatic;[54] however, this is only one of several competing theories.[4][49] About half of current scholarly hypotheses on Ongota's origins align it with Afroasiatic in some way.[55]
  • Robert Hetzron proposed that Beja is not part of Cushitic, but a separate branch.[56] The prevailing opinion, however, is that Beja is a branch of Cushitic.[57]
  • The extinct Meroitic language has been proposed to represent a branch of Afroasiatic.[58] Although an Afroasiatic connection is sometimes viewed as refuted, it continues to be defended by scholars such as Edward Lipiński.[59]
  • The
    Kujarge language is usually considered part of the Chadic languages;[60] however, Roger Blench has proposed that it may be a separate branch of Afroasiatic.[61][62]

Further subdivisions

Some proposed Afroasiatic subdivisions[a]
Fleming 1983 Ehret 1995 Bender 1997 Diakonoff 1988 Militarev 2005
  • Omotic
  • Afroasiatic
    • Semitic
    • Old East Africa Cushitic[b]
    • Erythraic
      • Cushitic[c]
      • Beja
      • Egyptian
      • Berber-Chadic
        • Berber
        • Chadic
  • Omotic
  • Erythraean
    • Cushitic
    • North Erythrean
      • Chadic
      • Boreafrasian
        • Egyptian
        • Berber
        • Semitic
  • Omotic
  • Chadic
  • Central Afroasiatic
    • Egyptian
    • Macro-Cushtic
      • Berber
      • Cushitic
      • Semitic
  • East–West Afrasian
    • Semitic
    • Cushitic
    • Berber (Berbero-Libyan)
  • North-South Afrasian
    • Egyptian
    • Chadic
    • Omotic
  • Cushomotic
    • Cushitic
    • Omotic
  • North Afrasian
    • Semitic
    • African North Afrasian
      • Egyptian
      • Chado-Berber
        • Chadic
        • Berber

There is no agreement on the relationships between and subgrouping of the different Afroasiatic branches.[58] Whereas Marcel Cohen (1947) claimed he saw no evidence for internal subgroupings, numerous other scholars have made proposals,[63] with Carsten Peust counting 27 as of 2012.[64]

Common trends in proposals as of 2019 include using common or lacking grammatical features to argue that Omotic was the first language to branch off, often followed by Chadic.

Igor Diakonoff.[70]

Responding to the above, Tom Güldemann criticizes attempts at finding subgroupings based on common or lacking morphology by arguing that the presence or absence of morphological features is not a useful way of discerning subgroupings in Afroasiatic, because it can not be excluded that families currently lacking certain features did not have them in the past; this also means that the presence of morphological features cannot be taken as defining a subgroup.[71] Peust notes that other factors that can obscure genetic relationships between languages include the poor state of present documentation and understanding of particular language families (historically with Egyptian, presently with Omotic).[72] Gene Gragg likewise argues that more needs to be known about Omotic still, and that Afroasiatic linguists have still not found convincing isoglosses on which to base genetic distinctions.[65]

One way of avoiding the problem of determining which features are original and which are inherited is to use a computational methodology such as lexicostatistics, with one of the earliest attempts being Fleming 1983.[69] This is also the method used by Alexander Militarev and Sergei Starostin to create a family tree.[73][74] Fleming (2006) was a more recent attempt by Fleming, with a different result from Militarev and Starostin.[72] Hezekiah Bacovcin and David Wilson argue that this methodology is invalid for discerning linguistic sub-relationship.[70] They note the method's inability to detect various strong commonalities even between well-studied branches of AA.[75]

Classification history

A relationship between Hebrew, Arabic, and Aramaic and the Berber languages was perceived as early as the 9th century CE by the Hebrew grammarian and physician

Ernst Renan named these languages, related to Semitic but not Semitic, "Hamitic," in 1860 Carl Lottner proposed that they belonged to a single language family, and in 1876 Friedrich Müller first described them as a "Hamito-Semitic" language family.[16] Müller assumed that there existed a distinct "Hamitic" branch of the family that consisted of Egyptian, Berber, and Cushitic.[77] He did not include the Chadic languages, though contemporary Egyptologist Karl Richard Lepsius argued for the relation of Hausa to the Berber languages.[78] Some scholars would continue to regard Hausa as related to the other Afroasiatic languages, but the idea was controversial: many scholars refused to admit that the largely unwritten, "Negroid" Chadic languages were in the same family as the "Caucasian" ancient civilizations of the Egyptians and Semites.[79][80]

Distribution of ethnic groups in Africa (Afroasiatic/Hamito-Semitic-speaking in yellow)

An important development in the history of Afroasiatic scholarship – and the history of African linguistics – was the creation of the "

Christian Bunsen, and linguist Christian Bleek.[81] This theory connected the "Hamites", the originators of Hamitic languages, with (supposedly culturally superior) "Caucasians", who were assumed to have migrated into Africa and intermixed with indigenous "Negroid" Africans in ancient times.[58] The "Hamitic theory" would serve as the basis for Carl Meinhof's highly influential classification of African languages in his 1912 book Die Sprache der Hamiten.[82] On one hand, the "Hamitic" classification was justified partially based on linguistic features: for example, Meinhof split the presently-understood Chadic family into "Hamito-Chadic", and an unrelated non-Hamitic "Chadic" based on which languages possessed grammatical gender.[80] On the other hand, the classification also relied on non-linguistic anthropological and culturally contingent features, such as skin color, hair type, and lifestyle.[83] Ultimately, Meinhof's classification of Hamitic proved to include languages from every presently-recognized language family within Africa.[78]

The first scholar to question the existence of "Hamitic languages" was Marcel Cohen in 1924,[19] with skepticism also expressed by A. Klingenheben and Dietrich Westermann during the 1920s and '30s.[58] However, Meinhof's "Hamitic" classification remained prevalent throughout the early 20th century until it was definitively disproven by Joseph Greenberg in the 1940s, based on racial and anthropological data.[80][84] Instead, Greenberg proposed an Afroasiatic family consisting of five branches: Berber, Chadic, Cushitic, Egyptian, and Semitic.[58][85][19] Reluctance among some scholars to recognize Chadic as a branch of Afroasiatic persisted as late as the 1980s.[80] In 1969, Harold Fleming proposed that a group of languages classified by Greenberg as Cushitic were in fact their own independent "Omotic" branch—a proposal that has been widely, if not universally, accepted.[58] These six branches now constitute an academic consensus on the genetic structure of the family.[86]

Greenberg relied on his own method of

genetic validity.[91][9]

Origin

Timeline

There is no consensus as to when Proto-Afroasiatic was spoken.[43] The absolute latest date for when Proto-Afroasiatic could have been extant is c. 4000 BCE, after which Egyptian and the Semitic languages are firmly attested. However, in all likelihood these languages began to diverge well before this hard boundary.[65] The estimations offered by scholars as to when Proto-Afroasiatic was spoken vary widely, ranging from 18,000 BCE to 8,000 BCE.[43] An estimate at the youngest end of this range still makes Afroasiatic the oldest proven language family.[6] Contrasting proposals of an early emergence, Tom Güldemann has argued that less time may have been required for the divergence than is usually assumed, as it is possible for a language to rapidly restructure due to areal contact, with the evolution of Chadic (and likely also Omotic) serving as pertinent examples.[92]

Location

Likewise, no consensus exists as to where proto-Afroasiatic originated.

Proto-Cushitic speakers with economic transformations in the Sahara dating c. 8,500 ago, as well as the speakers of the Proto-Zenati variety of the Berber languages with an expansion across the Maghreb in the 5th century CE.[95]

An origin somewhere on the African continent has broad scholarly support,

hunter-gatherers,[92] arguing that there is no evidence of words in Proto-Afroasiatic related to agriculture or animal husbandry.[97] Christopher Ehret, S.O. Y. Keita, and Paul Newman also argue that archaeology does not indicate a spread of migrating farmers into Africa, but rather a gradual incorporation of animal husbandry into indigenous foraging cultures.[99] Ehret, in a separate publication, argued that the two principles in linguistic approaches for determining the origin of languages which are the principles of fewest moves and greatest diversity had put “beyond reasonable doubt” that the language family “had originated in the Horn of Africa”.[100]

A significant minority of scholars supports an Asian origin of Afroasiatic,

Natufian Culture, arguing that the reconstructed lexicon of flora and fauna, as well as farming and pastoralist vocabulary indicates that Proto-AA must have been spoken in this area.[103][104] Scholar Jared Diamond and archaeologist Peter Bellwood have taken up Militarev's arguments as part of their general argument that the spread of linguistic macrofamilies (such as Indo-European, Bantu, and Austro-Asiatic) can be associated with the development of agriculture; they argue that there is clear archaeological support for farming spreading from the Levant into Africa via the Nile valley.[105]

Phonological characteristics

Speech sample in Shilha (Berber branch)
Speech sample in the Semitic
Neo-Aramaic language, a descendant of Old Aramaic
Speech sample in Somali (Cushitic branch)
Speech sample in Classical Arabic (Semitic branch)

Afroasiatic languages share a number of phonetic and phonological features.[106]

Syllable structure

Egyptian, Cushitic, Berber, Omotic, and most languages in the Semitic branch all require a syllable to begin with a consonant (with the exception of some grammatical prefixes).

Igor Diakonoff argues that this constraint goes back to Proto-Afroasiatic.[108] Some Chadic languages allow a syllable to begin with a vowel,[107] however in many Chadic languages verbs must begin with a consonant. In Cushitic and Chadic languages, a glottal stop or glottal fricative may be inserted to prevent a word from beginning with a vowel.[63] Typically, syllables only begin with a single consonant.[109]

With the exception of some Chadic languages, all Afroasiatic languages allow both closed and open syllables; many Chadic languages do not allow a syllable to end in a consonant.[110] Most words end in a vowel in Omotic and Cushitic, making syllable-final consonant clusters rare.[109] Diakonoff argues that proto-Afroasiatic syllables disallowed consonant clusters or vowels at the end of a syllable.[108]

Syllable weight plays an important role in AA, especially in Chadic; it can affect the form of affixes attached to a word.[111]

Consonant systems

Several Afroasiatic languages have large consonant inventories, and it is likely that this is inherited from proto-Afroasiatic.

Tigrinya, and some Chadic languages, there is no underlying phoneme [p] at all.[115]

Most, if not all branches of Afroasiatic distinguish between

implosive consonants in the different branches.[116] It is generally agreed that only the obstruents had a contrast between voiceless and voiced forms in Proto-Afroasiatic, whereas continuants were voiceless.[118]

A form of long-distance consonant

assimilation known as consonant harmony is attested in Berber, Chadic, Cushitic, and Semitic: it usually affects features such as pharyngealization, palatalization, and labialization.[119] Several Omotic languages have "sibilant harmony", meaning that all sibilants (s, sh, z, ts, etc.) in a word must match.[120]

Consonant incompatibility

Examples of root consonant incompatibilities from Egyptian, after Allen 2020a
consonant cannot occur with
p b, f, m, h
r ꜣ, b
h, ḥ, ẖ, q, k, g, ṯ, ḏ
s ḥ, z
t ꜥ, z, q, g, d, ḏ

Restrictions against the co-occurrence of certain, usually similar, consonants in verbal roots can be found in all Afroasiatic branches, though they are only weakly attested in Chadic and Omotic.

A set of constraints, developed originally by Joseph Greenberg on the basis of Arabic, has been claimed to be typical for Afroasiatic languages.[127] Greenberg divided Semitic consonants into four types: "back consonants" (glottal, pharyngeal, uvular, laryngeal, and velar consonants), "front consonants" (dental or alveolar consonants), liquid consonants, and labial consonants. He showed that, generally, any consonant from one of these groups could combine with consonants from any other group, but could not be used together with consonants from the same group.[124] Additionally, he showed that Proto-Semitic restricted a sequence of two identical consonants in the first and second position of the triliteral root.[128] These rules also have a number of exceptions:

  1. velar consonants can occur with pharyngeals or laryngeals;[129]
  2. dental consonants can co-occur with sibilants;[130] However, there are no Proto-Semitic verbal roots with and a sibilant, and roots with d and a sibilant are uncommon. In all attested cases of a dental and a sibilant, the sibilant occurs in first position and the dental in second.[129]

Similar exceptions can be demonstrated for the other AA branches that have these restrictions to their root formation.

James P. Allen has demonstrated that slightly different rules apply to Egyptian: for instance, Egyptian allows two identical consonants in some roots, and disallows velars from occurring with pharyngeals.[131]

Vowel systems

There is a large variety of vocalic systems in AA,[114] and attempts to reconstruct the vocalic system of Proto-Afroasiatic vary considerably.[118] All branches of Afroasiatic have a limited number of underlying vowels (between two and seven), but the number of phonetic vowels can be much larger. The quality of the underlying vowels varies considerably by language; the most common vowel throughout AA is schwa.[132] In the different languages, central vowels are often inserted to break up consonant clusters (a form of epenthesis).[114] Various Semitic, Cushitic, Berber, and Chadic languages, including Arabic, Amharic, Berber, Somali, and East Dangla, also exhibit various types of vowel harmony.[133]

Tones

The majority of AA languages are

pitch accent.[134]

Examples of tones marking lexical and morphological changes in some AA languages, after Frajzyngier 2012
Language Examples
Somali (Cushitic) díbi bull, absolutive case dibi bull, nominative case dibí bull, genitive case
ínan, boy inán girl
Bench (Omotic) k'áyts' work! do it! (active imperative) k'àyts' be done! (passive imperative)
Hausa (Chadic) màatáa woman, wife máatáa women, wives
dáfàa to cook (infinitive) dàfáa cook! (imperative)

Similarities in grammar, syntax, and morphology

At present, there is no generally accepted reconstruction of Proto-Afroasiatic grammar, syntax, or morphology, nor one for any of the sub-branches besides Egyptian. This means that it is difficult to know which features in Afroasiatic languages are retentions, and which are innovations.[137] Moreover, all Afroasiatic languages have long been in contact with other language families and with each other, leading to the possibility of widespread borrowing both within Afroasiatic and from unrelated languages.[138] There are nevertheless a number of commonly observed features in Afroasiatic morphology and derivation, including the use of suffixes, infixes, vowel lengthening and shortening as a morphological change, as well as the use of tone changes to indicate morphology.[139] Further commonalities and differences are explored in more detail below.

General features

Consonantal root structures

A widely attested feature in AA languages is a consonantal structure into which various vocalic "templates" are placed.[134] This structure is particularly visible in the verbs,[140] and is particularly noticeable in Semitic.[134] Besides for Semitic, vocalic templates are well attested for Cushitic and Berber,[141] where, along with Chadic, it is less productive; it is absent in Omotic.[134] For Egyptian, evidence for the root-and-template structure exists from Coptic.[141] In Semitic, Egyptian, Berber, verbs have no inherent vowels at all; the vowels found in a given stem are dependent on the vocalic template.[142] In Chadic, verb stems can include an inherent vowel as well.[63]

Most Semitic verbs are triliteral (have three consonants), whereas most Chadic, Omotic, and Cushitic verbs are biliteral (having two consonants).[143] The degree to which the Proto-AA verbal root was triliteral is debated.[92] It may have originally been mostly biconsonantal, to which various affixes (such as verbal extensions) were then added and lexicalized.[134] Although any root could theoretically be used to create a noun or a verb, there is evidence for the existence of distinct noun and verb roots, which behave in different ways.[144]

Examples of verbal templates in AA languages, after Gragg 2019
Language Akkadian (Semitic) Berber Beja (Cushitic) Ron/Daffo (Chadic) Coptic (Egyptian)
Root p-r-s to divide k-n-f to roast d-b-l to gather m-(w)-t to die k-t to build
Templates iprus- (preterite) ǎknəf (aorist) -dbil- (past) mot (perfective) kôt (infintive)
iparras- (present) əknǎf (perfective) -i:-dbil- (aorist) mwaát (imperfective) kêt (qualitative)
iptaras (perfect) əkǎnnǎf (imperfective) i:-dbil- (modal)
əknəf (neg. perfective) da:n-bi:l (present sg)
əkənnəf (neg. imperfective) -e:-dbil- (present pl)
-dabi:l- (negative)

As part of these templates, the alternation (apophony) between high vowels (e.g. i, u) and a low vowel (a) in verbal forms is usually described as one of the main characteristics of AA languages: this change codes a variety of different functions.[145] It is unclear whether this system is a common AA trait;[146] the Chadic examples, for instance, show signs of originally deriving from affixes, which could explain the origins of the alterations in other languages as well.[145]

Word order

It remains unclear what

subject-verb-object word order.[147]

Reduplication and gemination

Afroasiatic Languages use the processes of reduplication and gemination (which often overlap in meaning) to derive nouns, verbs, adjectives, and adverbs throughout the AA language family. Gemination in particular is one of the typical features of AA.[149] Full or partial reduplication of the verb is often used to derive forms showing repeated action (pluractionality), though it is unclear if this is an inherited feature or has been widely borrowed.[150]

Nouns

Grammatical gender and number

Use of T on feminine nouns, using data from Souag 2023
Kabyle (Berber) Hausa (Chadic) Beja (Cushitic) Egyptian Arabic (Semitic)
wəl-t 'daughter' yārinyà-r̃ 'the girl' (r̃ < final -t)
ʔo:(r)-t 'a daughter'
t-ʔo:r 'the daughter'
zꜣ-t 'daughter' bin-t 'daughter'

The assignment of nouns and pronouns to either masculine or feminine gender is present in all branches – but not all languages – of the Afroasiatic family.

singulative marker in some languages.[151]

Some examples of internal plurals in AA, using data from Gragg 2019 and Meyer & Wolff 2019
Language Meaning Singular Plural
Ge'ez (Semitic) king nɨgus nägäs-t
Teshelhiyt (Berber) country ta-mazir-t ti-mizar
Afar (Cushitic) body galab galo:b-a
Hausa (Chadic) stream gulbi gulà:be:
Mubi (Chadic) eye irin aràn

Afroasiatic languages have a variety of ways of marking plurals; in some branches, nouns change gender from singular to plural (gender polarity),[153] while in others, plural forms are ungendered.[155] In addition to marking plurals via a number of affixes (with the suffixes -*uu/-*w and -*n(a) widely attested), several AA languages make use of internal vowel change (apophony) and/or insertion (epenthesis).[156] These so-called "internal a" or "broken" plurals are securely attested in Semitic, Berber, Cushitic, and Chadic, although it is unclear if the Chadic examples are an independent development.[157][e] Another common method of forming plurals is reduplication.[159]

Noun cases and states

Subject-Object case marking in some AA branches, using data in Gragg 2019, Huehnergard 2011 and Bender 2000
Case Oromo (Cushitic) Berber Akkadian (Semitic) Wolaitta (Omotic)
Masculine Feminine Masculine Feminine Masculine Feminine Masculine Feminine
Nominative/bound nam-(n)i boy intal-t-i girl u-frux boy t-frux-t girl šarr-u-m king šarr-at-u-m queen keett-i house macci-yo woman
Accusative/absolutive/unbound nam-a intal-a a-frux t-a-frux-t šarr-a-m šarr-at-a-m keett-a macci-ya

Nouns cases are found in the Semitic, Berber, Cushitic, and Omotic branches. They are not found in Chadic languages, and there is no evidence for cases in Egyptian.[160] A common pattern in AA languages with case is for the nominative to be marked by -u or -i, and the accusative to be marked by -a.[161] However, the number and types of cases varies across AA and also within the individual branches.[160] Some languages in AA have a marked nominative alignment, a feature which may date back to Proto-Afroasiatic.[162][163] Zygmont Frajzyngier states that a general characteristic of case marking in AA languages is that it tends to mark roles such as genitive, dative, locative, etc. rather than the subject and object.[164]

Noun states in different AA branches, using data from Allen 2020, Lipiński 2001, Mous 2012, and Kossmann 2012
Language Free/absolute state Construct State Additional state
Aramaic (Semitic) malkā(h) queen malkat Emphatic: malkətā
Coptic (Egyptian) jôj head jaj- Pronominal: jô-
Iraqw (Cushitic) afee mouths afé-r
Riffian (Berber) a-ryaz man Annexed: wə-ryaz

A second category, which partially overlaps with case, is the AA linguistic category of "state." Linguists use the term "state" to refer to different things in different languages. In Cushitic and Semitic, nouns exist in the "free state" or the "construct state". The construct state is a special, usually reduced form of a noun, which is used when the noun is possessed by another noun (Semitic) or is modified by an adjective or relative clause (Cushitic). [165] Edward Lipiński refers to Semitic nouns as having four states: absolute (free/indeterminate), construct, determinate, and predicate.[166] Coptic and Egyptian grammar also refers to nouns having a "free" (absolute) state, a "construct state," and a "pronominal state." The construct state is used when a noun becomes unstressed as the first element of a compound, whereas the pronominal state is used when the noun has a suffixed possessive pronoun.[167][168] Berber instead contrasts between the "free state" and the "annexed state," the latter of which is used for a variety of purposes, including for subjects placed after a verb and after certain prepositions.[169]

Modifiers and agreement

There is no strict distinction between adjectives, nouns, and adverbs in Afroasaiatic.[170] All branches of Afroasiatic have a lexical category of adjectives except for Chadic;[159] some Chadic languages do have adjectives, however. In Berber languages, adjectives are rare and are mostly replaced by nouns of quality and stative verbs. [171] In different languages, adjectives (and other modifiers) must either precede or follow the noun.[172] In most AA languages, numerals precede the noun.[171]

In those languages that have adjectives, they can take gender and number markings, which, in some cases, agree with the gender and number of the noun they are modifying.[159] However, in Omotic, adjectives do not agree with nouns: sometimes, they only but take gender and number marking only when they are used as nouns, in other cases, they take gender and number marking only when they follow the noun (the noun then receives no marking).[173]

Masculine, Feminine, Plural agreement patterns in N T N, using data from Greenberg 1960[f]
Language meaning Masculine Feminine Plural[g]
Old South Arabian (Semitic) this ð-n ð-t ʔl-n
Egyptian this (p-n) t-n n-n
Beja (Cushitic) this be-n be-t bal-īn
Tuareg (Berber) relative verb form ilkəm-ən təlkəm-ət ilkəm-ən-in
Hausa (Chadic) possessive base na- ta- na-

A widespread pattern of gender and number marking in Afroasiatic, found on demonstratives, articles, adjectives, and relative markers, is a consonant N for masculine, T for feminine, and N for plural. This can be found in Semitic, Egyptian, Beja, Berber, and Chadic.[174][154] A system K (masculine), T (feminine), and H (plural) can be found in Cushitic, Chadic, with masculine K also appearing in Omotic.[118][116] The feminine marker T is one of the most consistent aspects across the different branches of AA.[153]

Verb forms

Tenses, aspects, and moods (TAMs)

There is no agreement about which

imperfective verbal aspects, which can be found in Cushitic, Berber, Semitic, most Chadic languages, and some Omotic languages.[176] The Egyptian verbal system diverges greatly from that found in the other branches.[150] Additionally, it is common in Afroasiatic languages for the present/imperfective form to be a derived (marked) form of the verb, whereas in most other languages and language families the present tense is the default form of the verb.[177] Another common trait across the family is the use of a suppletive imperative for verbs of motion.[178]

"Prefix conjugation"

Conjugation of verbs using prefixes that mark person, number, and gender can be found in Semitic, Berber, and in Cushitic,[150][179] where it is only found on a small set of frequent verbs.[180] These prefixes are clearly cognate across the branches, although their use within the verbal systems of the individual languages varies.[179] There is a general pattern in which n- is used for the first person plural, whereas t- is used for all forms of the second person regardless of plurality or gender, as well as feminine singular.[181] Prefixes of ʔ- (glottal stop) for the first person singular and y- for the third person masculine can also be reconstructed.[182] As there is no evidence for the "prefix conjugation" in Omotic, Chadic, or Egyptian, it is unclear whether this was a Proto-Afroasiatic feature that has been lost in those branches or is a shared innovation among Semitic, Berber, and Cushitic.[140][181][h]

The "prefix conjugation" in Afroasiatic, following Gragg 2019
Number Person Gender Akkadian (Semitic) Berber Beja (Cushitic)
Preterite Present Aorist Imperfective "Old Past" "Old Present" "New Present"
Singular 1 a-prus a-parras ăknəf-ăʕ əkănnăf-ăʕ ʔ-i:-dbíl ʔ-a-dbíl ʔ-a-danbí:l
2 m ta-prus ta-parras t-ăknəf-ət t-əkănnăf-ət t-i:-dbíl-a t-i-dbíl-a danbí:l-a
f ta-prus-i: ta-parras-i t-i:-dbíl-i t-i-dbíl-i danbí:l-i
3 m i-prus i-parras y-ăknəf y-əkănnăf ʔ-i:-dbíl ʔ-i-dbíl danbí:l
f ta-prus ta-parras t-ăknəf t-əkănnăf t-i:-dbíl t-i-dbíl
Plural 1 ni-prus ni-parras n-ăknəf n-əkănnăf n-i:-dbíl n-i-dbíl n-e:-dbíl
2 m ta-prus-a: a-parras t-ăknəf-ăm t-əkănnăf-ăm t-i:-dbíl-na t-i-dbíl-na t-e:-dbíl-na
f ta-parras t-ăknəf-măt t-əkănnăf-măt
3 m i-prus-u: ta-parras-i: ăknəf-ăn əkănnăf-ăn ʔ-i:-dbíl ʔ-i-dbíl ʔ-e:-dbíl-na
f i-prus-a: i-parras ăknəf-năt əkănnăf-năt

"Suffix conjugation"

Some AA branches have what is called a "suffix conjugation", formed by adding pronominal suffixes to indicate person, gender, and number to a verbal adjective.[150][184] In Akkadian, Egyptian, Berber, and Cushitic this forms a "stative conjugation", used to express the state or result of an action; the same endings as in Akkadian and Egyptian are also present in the West Semitic perfective verb form.[185][186] In Akkadian and Egyptian, the suffixes appear to be reduced forms of the independent pronouns (see Pronouns); the obvious correspondence between the endings in the two branches has been argued to show that Egyptian and Semitic are closely related.[187] While some scholars posit an AA origin for this form, it is possible that the Berber and Cushitic forms are independent developments,[150] as they show significant differences from the Egyptian and Semitic forms. The Cushitic forms in particular may be derived from morphology found in subordinate clauses.[179]

The "suffix conjugation" in Afroasiatic, following Gragg 2019
Number Person Gender Akkadian (Semitic) Egyptian Berber Afar (Cushitic)
Singular 1 pars-a:ku sḏm-kw măttit-ăʕ miʕ-iyo-h
2 m pars-a:ta sḏm-tj măttit-ət miʕ-ito-h
f pars-a:ti
3 m paris sḏm-w măttit meʕ-e-h
f pars-at sḏm-tj măttit-ăt
Plural 1 pars-a:nu sḏm-wjn măttit-it miʕ-ino-h
2 m pars-a:tunu sḏm-tjwnj miʕ-ito:nu-h
f pars-a:tina
3 m pars-u: sḏm-wj moʕ-o:nu-h
f pars-a:

Common derivational affixes

M-prefix noun derivation

A prefix in m- is the most widely attested affix in AA that is used to derive nouns,[153][188] and is one the features Joseph Greenberg used to diagnose membership in the family. It forms agent nouns, place nouns, and instrument nouns.[118][56] In some branches, it can also derive abstract nouns and participles.[188] Omotic, meanwhile, shows evidence for a non-productive prefix mV- associated with the feminine gender.[189] Christopher Ehret has argued that this prefix is a later development that was not present in Proto-Afro-Asiatic, but rather derived from a PAA indefinite pronoun *m-.[190] Such an etymology is rejected by A. Zaborski and Gábor Takács, the latter of whom argues for a PAA *ma- that unites all or some of the meanings in the modern languages.[191]

Examples of m-prefix noun derivations, using data from Meyer & Wolff 2019, Beylage 2018, and Wilson 2020
Language Root Agent/Instrument Place/Abstract
Egyptian swr to drink m-swr drinking bowl
Arabic (Semitic) k-t-b to write mu-katib-un writer ma-ktab-un school
Hausa (Chadic) hayf- to give birth má-hàif-íi father má-háif-áa birthplace
Beja (Cushitic) firi to give birth mi-frey birth
Tuareg (Berber) äks to eat em-äks eater

Verbal extensions

Common verbal extensions in Afroasiatic, using data from Wilson 2020, Bubenik 2023, and Kossmann 2007
Language Causative *s- Reflexive/middle *t- Passive *n-
Akkadian (Semitic) u-š-apris 'make cut' mi-t-gurum 'agree (with one another)' i-p-paris (> *i-n-paris) 'be cut'
Figuig (Berber) ssu-fəɣ 'let out' i-ttə-ska 'it has been built' mmu-bḍa 'divide oneself'
Beja (Cushitic) s-dabil 'make gather' t-dabil 'be gathered' m-dabaal 'gather each other'
Egyptian s-ꜥnḫ 'make live' pr-tj 'is sent forth'[i] n-hp 'escape'[j]

Many AA languages use prefixes or suffixes (verbal extensions) to encode various pieces of information about the verb.

passive';[184] the prefixes appear with various related meanings in the individual daughter languages and branches.[195] Christopher Ehret has proposed that Proto-Afroasiatic originally had as many as thirty-seven separate verbal extensions, many of which then became fossilized as third consonants.[196] This theory has been criticized by some, such as Andrzej Zaborski and Alan Kaye, as being too many extensions to be realistic, though Zygmont Frajzyngier and Erin Shay note that some Chadic languages have as many as twelve extensions.[93]

"Nisba" derivation

The so-called "Nisba" is a suffix used to derive adjectives from nouns and, in Egyptian, also from prepositions.[197] It is found in Egyptian, Semitic, and possibly, in some relic forms, Berber.[198] The suffix has the same basic form in Egyptian and Semitic,[197] taking the form -i(y) in Semitic and being written -j in Egyptian. The Semitic and Cushitic genitive case in -i/-ii may be related to "nisba" adjective derivation.[199][158]

"Nisba" derivation in Semitic and Egyptian, using data from Wilson 2020 and Beylage 2018
Language Noun/preposition Derived adjective
Hebrew (Semitic) yārēaḥ moon yərēḥī lunar
Egyptian nṯr god nṯr.j divine
ḥr upon ḥr.j upper, which is upon

Due to its presence in the oldest attested and best-known AA branches, nisba derivation is often thought of as a "quintessentially Afroasiatic feature".[158][200] Christopher Ehret argues for its presence in Proto-Afroasiatic and for its attestation in some form in all branches, with a shape -*ay in addition to -*iy in some cases.[201]

Vocabulary comparison

Pronouns

The forms of the pronouns are very stable throughout Afroasiatic (excluding Omotic),

sibilant consonant (s, sh).[203] Other commonalities are masculine and feminine forms used in both the second and third persons, except in Cushitic and Omotic.[147] These pronouns tend to show a masculine "u" and a feminine "i".[116] The Omotic forms of the personal pronouns differ from the others, with only the plural forms in North Omotic appearing potentially to be cognate.[204]

Pronouns in the Afroasiatic family, following Gragg 2019
Meaning North Omotic (Yemsa) Beja Cushitic (Baniamer) East Cushitic (Somali) West Chadic (Hausa) East Chadic (Mubi) Egyptian East Semitic (Akkadian) West Semitic (Arabic) Berber (Tashelhiyt)
'I' (ind.) aní aní-ga ni: ndé jnk ana:ku ʔana nkki
'me, my' (dep.) -ná-
-tá-
-u: -ʔe na -j
wj
-i:
-ya
-i:
-ni:
-i
'we' (ind.) ìnno hinín anná-ga
inná-ga
mu: ána
éné
jnn ni:nu: naħnu nkkwni
'you' (masc. sing. ind.) barú:k adí-ga kai kám nt-k at-ta ʔan-ta kiji
'you' (fem. sing. ind.) batú:k ke: kín nt-ṯ at-ti ʔan-ti kmmi (f)
'you' (masc. sing., dep.) -né- -ú:k(a) ku ka -k -ka -ka -k
'you' (fem. sing., dep.) -ú:k(i) ku ki -ṯ -ki -ki -m
'you' (plural, dep.) -nitì- -ú:kna idin ku ká(n) -ṯn -kunu (m)
-kina (f)
-kum (m)
-kunna (f)
-un (m)
-un-t (f)
'he' (ind.) bár barú:s isá-ga ši: ár nt-f šu ntta (m)
'she' (ind.) batú:s ijá-ga ita tír nt-s ši hiya ntta-t
'he' (dep.) -bá- -ūs ši à -f
sw
-šu -hu -s
'she' (dep.) ta -s
sy
-ša -ha:

Numerals

Unlike in the

Indo-European or Austronesian language families, numerals in AA languages cannot be traced to a proto-system.[205][206] The Cushitic and Chadic numeral systems appear to have originally been base 5. The system in Berber, Egyptian, and Semitic, however, has independent words for the numbers 6–9.[207] Thus, it is possible that the numerals in Egyptian, Berber, and Semitic are more closely related, whereas the Cushitic and Chadic numerals are more closely related to each other.[208][209] Modern Chadic numeral systems are sometimes decimal, having separate names for the numbers 1–10, and sometimes base-5, deriving the numbers 6–9 from the numbers 1–5 in some way.[210] Some families show more than one word for a numeral: Chadic, Semitic, and Berber each have two words for two,[211][212] and Semitic has four words for one.[213] Andrzej Zaborski further notes that the numbers "one", "two", and "five" are particularly susceptible to replacement by new words, with "five" often based on a word meaning "hand".[205]

Another factor making comparisons of AA numeral systems difficult is the possibility of

Numerals from throughout Afroasiatic, using data from Blažek 2017, Blažek 2018, Lipiński 2001, and Frajzyngier & Shay 2012a
Meaning Egyptian Tuareg (Berber) Akkadian (East Semitic) Arabic (West Semitic) Beja (North Cushitic) West Central Oromo (Cushitic) Lele (East Chadic) Gidar (Central Chadic) Bench (North Omotic)
One m. wꜥ yiwən, yan, iğ ištēn wāḥid gáal tokko pínà tákà mat'
f. wꜥ.t yiwət, išt ištiāt wāḥida gáat
Two m. sn.wj sin, sən šinā ʔiṯnāni máloob lama súlà nam
f. sn.tj snat, sənt šittā ʔiṯnatāni máloot
Three m. ḫmt.w ḵraḍ, šaṛḍ šalāšat ṯalāṯa mháy sadii súbù hókù kaz
f. ḫmt.t ḵraṭt, šaṛṭ šalāš ṯalāṯ mháyt
Four m. (j)fd.w kkuẓ erbet(t) ʔarbaʕa faḍíg afur pórìn póɗó od
f. (j)fd.t kkuẓt erba ʔarbaʕ faḍígt
Five m. dj.w səmmus, afus ḫamšat ḫamsa áy šani bày ɬé ut͡ʃ
f. dj.t səmmust ḫamiš ḫams áyt
Six m sjs.w sḍis šiššet sitta aságwir jaha ménéŋ ɬré sapm
f. sjs.t sḍist šiš(š) sitt asagwitt
Seven m sfḫ.w sa sebet(t) sabʕa asarámaab tolba mátàlíŋ bùhúl napm
f. sfḫ.t sat seba sabʕ asarámaat
Eight m. ḫmn.w tam samānat ṯamāniya asúmhay saddet jurgù dòdòpórò nyartn
f. ḫmn.t tamt samānē ṯamānin asúmhayt
Nine m. psḏ.w tẓa tišīt tisʕa aššaḍíg sagal célà váyták irstn
f. psḏ.t tẓat tiše tisʕ aššaḍígt
Ten m. mḏ.w mraw ešeret ʕašara támin kuḍan gòrò kláù tam
f. mḏ.t mrawt ešer ʕašr támint

Cognates

Afroasiatic languages share a vocabulary of Proto-Afroasiatic origin to varying extents.[217] Writing in 2004, John Huehnergard notes the great difficulty in establishing cognate sets across the family.[184] Identifying cognates is difficult because the languages in question are often separated by thousands of years of development and many languages within the family have long been in contact with each other, raising the possibility of loanwords.[7] Work is also hampered because of the poor state of documentation of many languages.[218]

There are two etymological dictionaries of Afroasiatic, one by Christopher Ehret, and one by Vladimir Orel and Olga Stolbova, both from 1995. Both works provide highly divergent reconstructions and have been heavily criticized by other scholars.

sound correspondences.[112]

Some widely recognized cognates in Afroasiatic, following Hayward 2000, Gragg 2019, and Huehnergard 2004[k]
Meaning Proto-Afroasiatic Omotic Cushitic Chadic Egyptian Semitic Berber
Ehret 1995[l] Orel & Stolbova 1995
to strike, to squeeze *bak- Gamo bak- 'strike' Afar bak Wandala bak 'to strike, beat' bk 'kill (with a sword)' Arabic bkk 'to squeeze, tear' Tuareg bakkat 'to strike, pound'
blood *dîm-
*dâm-
*dam- Kaffa damo 'blood';
Aari zomʔi 'to blood'
(cf. Oromo di:ma 'red') Bolewa dom (cf. jdmj 'red linen') Akkadian damu 'blood' Ghadames dəmmm-ən 'blood'
food *kamaʔ- / *kamay- Afar okm- 'to eat' Hausa ka:ma:ma: 'snack';
Tumak ka:m 'mush'
kmj 'food'
to be old, elder *gâd-/gûd- *gad- Oromo gada 'age group, generation';
Burji gad-uwa 'old man'
Ngizim gad'e 'old' Arabic gadd- 'grandfather, ancestor'
to say *geh- *gay- Sheko ge 'to say';
Aari gai- 'to say'
Hausa gaya 'to say' ḏwj 'to call, say' (cf. Hebrew gʕy 'to shout')
tongue *lis'- 'to lick' *les- 'tongue' Kaffa mi-laso 'tongue' Mwaghavul liis tongue,
Gisiga eles 'tongue
ns 'tongue' Akkadian liša:nu 'tongue' Kabyle iləs 'tongue'
to die *maaw- *mawut- Rendille amut 'to die, to be ill' Hausa mutu 'to die',
Mubi ma:t 'to die'
mwt 'to die' Hebrew mwt, 'to die'
Ge'ez mo:ta 'to die'
Kabyle ammat 'to die'
to fly, to soar *pîr- *pir- (cf. Yemsa fill- 'to jump';
Dime far 'to jump')
Beja fir 'to fly' Hausa fi:ra 'to soar';
Mafa parr, perr 'bird's flight'
pꜣ 'to fly';
prj 'to soar, rise'
Ugaritic pr 'to flee';
Arabic frr 'to flee'
Ahogar fərə-t 'to fly'
name *sǔm / *sǐm- *süm- Hausa su:na: 'name';
Sura sun 'name';
Ga'anda ɬim 'name'
Akkadian šumu 'name'
to sour *s'ăm- Mocha č'àm- 'to be bitter' PEC *cam- 'to rot' *s'am 'sour' smj 'curds' Arabic sumūț 'to begin to turn sour'
to spit *tuf- *tuf- Beja tuf 'to spit';
Kemant təff y- 'to spit';
Somali tuf 'to spit'
tf 'to spit' Aramaic tpp 'to spit';
Arabic tff 'to spit'
to rend, tear *zaaʕ- Gamo zaʔ 'to rend, split' Dahalo ḏaaʕ- 'to rend, to tear (of an animal tearing its prey)'
Kw'adza daʔ- 'to bite'[m]
Ngizim dáar- 'to cut into long strips' Arabic zaʕy- 'to snatch violently from, tear out'
Abbreviations: PEC='Proto-Eastern Cushtic'.

See also

Notes

  1. ^ In this display, the universally recognized primary branches of AA are in bold. Individual languages designated as primary branches are in italics.
  2. ^ Old East Africa Cushitic =South Cushitic and Yaaku.
  3. ^ Fleming excludes South Cushitic, Yaaku, and Beja from Cushitic proper.
  4. ^ Some scholars reconstruct "emphatic" consonants for Egyptian and some do not.[117]
  5. ^ As Egyptian is spelled without vowels, it is difficult to know whether it had internal change plurals. There is some evidence from Coptic, but this may be unrelated to AA. There is also some evidence from cuneiform transcriptions of Egyptian words.[158]
  6. ^ Forms that do not follow the pattern are in parentheses.
  7. ^ Some languages have an additional feminine plural form in -t that is not listed here.
  8. ^ Traditionally, the Hausa subject pronouns have been compared to the prefix conjugation. However, since the 1970s and '80s, comparisons of other Chadic subject pronouns with the Hausa ones have convinced most scholars that the similarity to the prefix conjugation is incidental.[183]
  9. ^ The Egyptian passive suffix is solely morphological and does not form a unique stem.[192]
  10. ^ The Egyptian prefix has a middle voice/intransitive/or passive meaning.[193]
  11. ^ Many of these roots have other proposed cognates that are not included on the table.
  12. comparison of related languages
    .
  13. ^ Ehret's dictionary lists "Proto-South-Cushitic *daaʕ- 'to rend, tear', a root he reconstructs from the two words listed here in Ehret 1980.

Citations

  1. ^ Sands 2009, pp. 559–580.
  2. ^ a b c d Almansa-Villatoro & Štubňová Nigrelli 2023, p. 3.
  3. ^ a b Eberhard, Simons & Fennig 2021, Summary by language family.
  4. ^ a b c d e Sands 2009, p. 565.
  5. ^ a b Meyer & Wolff 2019, p. 249.
  6. ^ a b Nichols 2003, p. 300.
  7. ^ a b c Gragg 2019, p. 41.
  8. ^ a b Güldemann 2018, p. 315-316.
  9. ^ a b c d Frajzyngier & Shay 2012, p. 3.
  10. ^ a b c Zaborski 2011.
  11. ^ Hetzron 2009, p. 454.
  12. ^ a b c d e Porkhomovsky 2020, p. 270.
  13. ^ Lipiński 2001, pp. 21–22.
  14. ^ Meyer & Wolff 2019, p. 250.
  15. ^ Porkhomovsky 2020, pp. 269–270.
  16. ^ a b c Porkhomovsky 2020, p. 269.
  17. ^ Solleveld 2020, p. 204.
  18. ^ a b Huehnergard 2004, p. 138.
  19. ^ a b c Frajzyngier & Shay 2012, p. 5.
  20. ^ a b Dimmendaal 2008, p. 840.
  21. ^ a b Hetzron 2009, p. 545.
  22. ^ a b Almansa-Villatoro & Štubňová Nigrelli 2023, p. 4.
  23. ^ Huehnergard 2004, p. 138-139.
  24. ^ a b Frajzyngier & Shay 2012, p. 1.
  25. ^ Gragg 2019, pp. 23–24.
  26. ^ Meyer & Wolff 2019, p. 252-253.
  27. ^ Güldemann 2018, p. 324.
  28. ^ Lipiński 2001, p. 34.
  29. ^ Meyer & Wolff 2019, p. 253.
  30. ^ a b c Gragg 2019, p. 24.
  31. ^ a b Güldemann 2018, p. 325.
  32. ^ Lipiński 2001, p. 37.
  33. ^ a b c Meyer & Wolff 2019, p. 254.
  34. ^ a b c Gragg 2019, p. 29.
  35. ^ a b c Gragg 2019, p. 27.
  36. ^ Lipiński 2001, p. 39.
  37. ^ Güldemann 2018, pp. 342–343.
  38. ^ a b c d e Meyer & Wolff 2019, p. 259.
  39. ^ a b c d Appleyard 2012, p. 39.
  40. ^ Gragg 2019, pp. 25–26.
  41. ^ Lipiński 2001, p. 30.
  42. ^ Allen 2013, p. 2.
  43. ^ a b c d e f Meyer & Wolff 2019, p. 252.
  44. ^ Allen 2013, pp. 4–5.
  45. ^ Allen 2013, p. 4.
  46. ^ Güldemann 2018, p. 322.
  47. ^ Frajzyngier & Shay 2012, p. 6.
  48. ^ Meyer & Wolff 2019, p. 261.
  49. ^ a b Huehnergard 2004, p. 139.
  50. ^ Gragg 2019, p. 26.
  51. ^ a b c Meyer & Wolff 2019, p. 262.
  52. ^ a b Lipiński 2001, p. 24.
  53. ^ Hayward 2000, pp. 78–80.
  54. ^ Fleming 2006.
  55. ^ Güldemann 2018, p. 342.
  56. ^ a b Huehnergard 2004, p. 140.
  57. ^ Güldemann 2018, p. 327.
  58. ^ a b c d e f Meyer & Wolff 2019, p. 251.
  59. ^ Güldemann 2018, p. 282.
  60. ^ Meyer & Wolff 2019, p. 258.
  61. ^ Peust 2012, p. 231.
  62. ^ Blench 2008.
  63. ^ a b c Frajzyngier 2018.
  64. ^ Peust 2012, p. 225-227.
  65. ^ a b c d e Gragg 2019, p. 43.
  66. ^ Blench 2006, p. 145.
  67. ^ a b Sanker 2023, p. 29.
  68. ^ Güldemann 2018, pp. 312–313.
  69. ^ a b c d Blench 2006, p. 144.
  70. ^ a b Bacovcin & Wilson 2018, p. 422.
  71. ^ Güldemann 2018, p. 310.
  72. ^ a b Peust 2012, p. 227.
  73. ^ Militarev 2005, pp. 398–399.
  74. ^ Blažek 2013, p. 1.
  75. ^ Bacovcin & Wilson 2018, p. 427.
  76. ^ a b Lipiński 2001, p. 21.
  77. ^ a b Frajzyngier & Shay 2012, p. 4.
  78. ^ a b Hayward 2000, p. 84.
  79. ^ Ruhlen 1991, pp. 87–88.
  80. ^ a b c d Porkhomovsky 2020, p. 271.
  81. ^ Solleveld 2020, p. 204-205.
  82. ^ Güldemann 2018, p. 61.
  83. ^ Ruhlen 1991, pp. 82–83.
  84. ^ Güldemann 2018, p. 309.
  85. ^ Gragg 2019, p. 22.
  86. ^ Porkhomovsky 2020, p. 272.
  87. ^ Hayward 2000, pp. 86–87.
  88. ^ Hodge 1971, p. 11.
  89. ^ Winand 2023, p. 40.
  90. ^ Huehnergard 2023, p. 140.
  91. ^ Güldemann 2018, p. 347.
  92. ^ a b c Güldemann 2018, p. 311.
  93. ^ a b c Frajzyngier & Shay 2012, p. 13.
  94. ^ Blench 2006, pp. 148–150.
  95. ^ Ehret et al. 2023, p. 270.
  96. ^ a b Frajzyngier 2012, p. 13.
  97. ^ a b Ehret, Keita & Newman 2004, p. 1680.
  98. ^ Starostin 2017, p. 226.
  99. ^ Ehret, Keita & Newman 2004, pp. 1680–1681.
  100. .
  101. ^ Blench 2006, p. 150.
  102. ^ Almansa-Villatoro & Štubňová Nigrelli 2023, p. 5.
  103. ^ Militarev 2002.
  104. ^ Militarev 2009.
  105. ^ Diamond & Bellwood 2003, p. 601.
  106. ^ Meyer & Wolff 2019, p. 263.
  107. ^ a b Frajzyngier 2012, p. 511.
  108. ^ a b Diakonoff 1988, p. 42.
  109. ^ a b Meyer & Wolff 2019, p. 304.
  110. ^ Frajzyngier 2012, p. 511-512.
  111. ^ Frajzyngier 2012, p. 512.
  112. ^ a b Hayward 2000, p. 94.
  113. ^ a b Frajzyngier 2012, p. 508.
  114. ^ a b c d e Meyer & Wolff 2019, p. 264.
  115. ^ Frajzyngier 2012, p. 509.
  116. ^ a b c d Hetzron 2009, p. 548.
  117. ^ Huehnergard 2023, p. 142.
  118. ^ a b c d Frajzyngier & Shay 2012, p. 10.
  119. ^ Frajzyngier 2012, p. 517.
  120. ^ Amha 2012, pp. 439–440.
  121. ^ Bender 1978, p. 9-10.
  122. ^ Ehret 1995, p. 77, 488.
  123. ^ Ehret 1995, p. 395.
  124. ^ a b Greenberg 1950a, p. 178.
  125. ^ Edzard 1992, p. 153-154.
  126. ^ Takács 1999, pp. 323–332.
  127. ^ Vernet 2011, p. 1.
  128. ^ Greenberg 1950a, pp. 167–168.
  129. ^ a b Vernet 2011, p. 7.
  130. ^ a b Bender 1978, p. 10.
  131. ^ Allen 2020a, p. 90-92.
  132. ^ Frajzyngier 2012, p. 621.
  133. ^ Frajzyngier 2012, pp. 517–519.
  134. ^ a b c d e f Meyer & Wolff 2019, p. 265.
  135. ^ Frajzyngier 2012, p. 12.
  136. ^ Frajzyngier 2012, p. 513.
  137. ^ Frajzyngier & Shay 2020, p. 572-573.
  138. ^ Shay 2014, p. 574.
  139. ^ Frajzyngier 2012, pp. 529–530.
  140. ^ a b Gragg 2019, p. 36.
  141. ^ a b Gragg 2019, p. 37.
  142. ^ Frajzyngier 2012, p. 622.
  143. ^ Hayward 2000, p. 93.
  144. ^ Shay 2014, p. 576.
  145. ^ a b Frajzyngier 2012, p. 593.
  146. ^ Gragg 2019, p. 38.
  147. ^ a b c d Meyer & Wolff 2019, p. 269.
  148. ^ a b Güldemann 2018, p. 319.
  149. ^ Frajzyngier 2012, p. 531.
  150. ^ a b c d e Meyer & Wolff 2019, p. 268.
  151. ^ a b Frajzyngier 2012, p. 522.
  152. ^ Frajzyngier 2012, p. 523.
  153. ^ a b c d Meyer & Wolff 2019, p. 266.
  154. ^ a b Souag 2023, p. 308.
  155. ^ Frajzyngier 2012, p. 253.
  156. ^ Meyer & Wolff 2019, pp. 266–267.
  157. ^ Gragg 2019, pp. 40–41.
  158. ^ a b c Peust 2012, p. 243.
  159. ^ a b c Frajzyngier 2012, p. 538.
  160. ^ a b Frajzyngier 2012, p. 534.
  161. ^ Meyer & Wolff 2019, p. 319.
  162. ^ Frajzyngier 2012, pp. 535–536.
  163. ^ Güldemann 2018, p. 317.
  164. ^ Frajzyngier 2012, p. 535.
  165. ^ Frajzyngier 2012, p. 533.
  166. ^ Lipiński 2001, p. 265.
  167. ^ Allen 2020, p. 13.
  168. ^ Beylage 2018, p. 59.
  169. ^ Frajzyngier 2012, pp. 533–534.
  170. ^ Meyer & Wolff 2019, p. 312.
  171. ^ a b Meyer & Wolff 2019, p. 279.
  172. ^ Meyer & Wolff 2019, p. 298.
  173. ^ Frajzyngier 2012, pp. 539–540.
  174. ^ Güldemann 2018, p. 316.
  175. ^ Meyer & Wolff 2019, p. 267.
  176. ^ Frajzyngier 2012, p. 592.
  177. ^ Kouwenberg 2010, p. 90.
  178. ^ Frajzyngier & Shay 2020.
  179. ^ a b c Gragg 2019, p. 34.
  180. ^ Mous 2012, p. 391.
  181. ^ a b Güldemann 2018, p. 315.
  182. ^ Huehnergard 2004, p. 140-141.
  183. ^ Peust 2012, pp. 238–239.
  184. ^ a b c d Huehnergard 2004, p. 141.
  185. ^ Gragg 2019, p. 33.
  186. ^ Lipiński 2001, p. 360.
  187. ^ Gragg 2019, p. 33-24.
  188. ^ a b Wilson 2020, p. 123.
  189. ^ Shay 2014, p. 577.
  190. ^ Ehret 1995, p. 52.
  191. ^ Takács 2008, p. 8.
  192. ^ Stauder 2023, pp. 88–90.
  193. ^ Allen 2013, p. 94.
  194. ^ Frajzyngier 2012, p. 525.
  195. ^ Stauder 2023, p. 87.
  196. ^ Ehret 1995, pp. 27–34.
  197. ^ a b Beylage 2018, p. 115.
  198. ^ Wilson 2020, p. 47.
  199. ^ Huehnergard 2004, p. 148.
  200. ^ Wilson 2020, p. 168.
  201. ^ Ehret 1995, p. 16.
  202. ^ Frajzyngier 2012, pp. 523–524.
  203. ^ Güldemann 2018, p. 314-315.
  204. ^ Gragg 2019, p. 32.
  205. ^ a b Zaborski 1987, p. 317.
  206. ^ a b Kaye & Daniels 1992, p. 439.
  207. ^ Zaborski 1987, pp. 317–318.
  208. ^ Lipiński 2001, p. 280.
  209. ^ Kaye & Daniels 1992, pp. 440–441.
  210. ^ Frajzyngier & Shay 2012a, p. 273.
  211. ^ Kaye & Daniels 1992, p. 440.
  212. ^ Lipiński 2001, p. 284.
  213. ^ Lipiński 2001, p. 281.
  214. ^ Meyer & Wolff 2019, p. 295.
  215. ^ Lipiński 2001, p. 280-281.
  216. ^ Zaborski 1987, p. 325.
  217. ^ Meyer & Wolff 2019, p. 248.
  218. ^ Porkhomovsky 2020, p. 273.
  219. ^ Güldemann 2018, p. 317-318.
  220. ^ Güldemann 2018, p. 318.

Works cited

External links