Cyprinus carpio Linnaeus, 1758

Common Name: Common Carp

Synonyms and Other Names:

European carp, German carp, mirror carp, leather carp



Copyright Info

Identification: Adults are light gold to dark brown in color, with reddish fins. Scales cover the entire body and barbels accent the mouth (one per side). Head is triangular with a blunt snout and thick nose plate. Does not possess a true spine. Wheeler (1978); Becker (1983); Page and Burr (1991); Etnier and Starnes (1993); Jenkins and Burkhead (1994); Balon (1995). In Eurasia there are two poorly defined subspecies C. c. carpio and C. c. haematopterus; unfortunately, feral Common Carp, descendants of earlier escapees or introductions, have greatly confused the picture (Balon 1995). Several genetic strains—some bred in aquaculture or used as ornamentals (e.g., leather carp, mirror carp, Israeli carp, koi)—are recognized by some as separate varieties (Robison and Buchanan 1988; Balon 1995).


Size: 122 cm


Native Range: Eurasia (Page and Burr 1991; Balon 1995). Balon (1995) found that Cyprinus carpio evolved in the Caspian Sea, then migrated naturally to the Black and Aral Seas, east to eastern mainland Asia and west as far as the Danube River.


Table 1. Great Lakes region nonindigenous occurrences, the earliest and latest observations in each state/province, and the tally and names of HUCs with observations†. Names and dates are hyperlinked to their relevant specimen records. The list of references for all nonindigenous occurrences of Cyprinus carpio are found here.

Full list of USGS occurrences

State/ProvinceFirst ObservedLast ObservedTotal HUCs with observations†HUCs with observations†
18792000*
IL189420223Lake Michigan; Little Calumet-Galien; Pike-Root
IN194120203Little Calumet-Galien; St. Joseph; St. Joseph
MI1880202347Au Gres-Rifle; Au Sable; Betsie-Platte; Birch-Willow; Black-Macatawa; Boardman-Charlevoix; Cass; Cheboygan; Clinton; Detroit; Fishdam-Sturgeon; Flint; Great Lakes Region; Huron; Kalamazoo; Kawkawlin-Pine; Keweenaw Peninsula; Lake Erie; Lake Huron; Lake Michigan; Lake St. Clair; Lake Superior; Lone Lake-Ocqueoc; Lower Grand; Manistee; Maple; Menominee; Millecoquins Lake-Brevoort River; Muskegon; Ottawa-Stony; Pere Marquette-White; Pigeon-Wiscoggin; Pine; Raisin; Saginaw; Shiawassee; St. Clair; St. Joseph; St. Joseph; St. Marys; Sturgeon; Tacoosh-Whitefish; Thornapple; Thunder Bay; Tiffin; Tittabawassee; Upper Grand
MN196520152Lake Superior; St. Louis
NY1905202327Black; Buffalo-Eighteenmile; Chateaugay-English; Chaumont-Perch; Eastern Lake Erie; Grass; Headwaters St. Lawrence River; Indian; Irondequoit-Ninemile; Lake Champlain; Lake Erie; Lake Ontario; Lower Genesee; Mettawee River; Niagara River; Oak Orchard-Twelvemile; Oneida; Oswegatchie; Oswego; Raisin River-St. Lawrence River; Raquette; Salmon; Salmon-Sandy; Seneca; Southwestern Lake Ontario; St. Regis; Upper Genesee
OH1942202014Ashtabula-Chagrin; Auglaize; Blanchard; Cedar-Portage; Chautauqua-Conneaut; Cuyahoga; Lake Erie; Lower Maumee; Ottawa-Stony; Sandusky; Southern Lake Erie; St. Joseph; Upper Maumee; Western Lake Erie
ONT20182023*
PA198220142Chautauqua-Conneaut; Lake Erie
VT198020204Lake Champlain; Missiquoi River; Otter Creek; Richelieu River
WI1902202218Beartrap-Nemadji; Black-Presque Isle; Brule; Door-Kewaunee; Duck-Pensaukee; Lake Michigan; Lake Superior; Lake Winnebago; Lower Fox; Manitowoc-Sheboygan; Menominee; Milwaukee; Oconto; Peshtigo; Pike-Root; St. Louis; Upper Fox; Wolf

Table last updated 4/23/2024

† Populations may not be currently present.

* HUCs are not listed for areas where the observation(s) cannot be approximated to a HUC (e.g. state centroids or Canadian provinces).


Ecology: The species generally inhabits lakes, ponds, and the lower sections of rivers (usually with moderately flowing or standing water), but is also known from brackish-water estuaries, backwaters, and bays (Barus et al. 2001). In its native range, the species occurs in coastal areas of the Caspian and Aral Seas (Berg 1964; Barus et al. 2001) as well as the estuaries of large Ukrainian and Russian rivers. Crivelli (1981) reported that the Common Carp occurred in brackish-water marshes with salinities up to 14 ppt in southern France. In North America, the Common Carp inhabits brackish and saline coastal waters of several states bordering the Atlantic and Pacific Oceans and Gulf of Mexico (Schwartz 1964; Moyle 2002) as well as the Atlantic and Pacific coasts of Canada (McCrimmon 1968). It has been captured in U.S. waters with salinities as high as 17.6 ppt (Schwartz 1964). In the U.S., the Common Carp is more abundant in manmade impoundments, lakes, and turbid sluggish streams receiving sewage or agricultural runoff, and less abundant in clear waters or streams with a high gradient (Pflieger 1975; Trautman 1981; Ross 2001; Boschung and Mayden 2004; Ramirez-Herrejon et al. 2015). Pflieger (1975) noted that the Common Carp tends to concentrate in large numbers where cannery or slaughter-house wastes are emptied into streams.

Larval Common Carp feed primarily on zooplankton. In its native range, juveniles and adults feed on benthic organisms (e.g., chironomids, gastropods and other larval insects), vegetation, detritus and plankton (e.g., cladocerans, copepods, amphipods, mysids). Feeding habits are similar in the U.S., where the diet is composed of organic detritus (primarily of plant origin), chironomids, small crustaceans, and gastropods (Summerfelt et al. 1971; Eder and Carlson 1977; Panek 1987). Common Carp have shown to be an important seed dispersal vector for aquatic plants (VonBank et al. 2018). Common Carp is very active when feeding and its movements often disturb sediments and increase turbidity, causing serious problems in some regions especially where the species is abundant. The species also retards the growth of submerged aquatic vegetation by feeding and uprooting plants (King and Hunt 1967). Silt resuspension and uprooting of aquatic plants caused by feeding activities can disturb spawning and nursery areas of native fishes (Ross 2001) as well as disrupt feeding of sight-oriented predators, such as bass and sunfish (Panek 1987).

In temperate regions, Common Carp spawn in spring when flows are high. Spawning occurs in submerged vegetation, plant debris, stones, and in artificial substrate. In more tropical climates Common Carp can spawn year round (Adámek et al. 2015). Females reach sexual maturity after 3–5 years and scatter sticky unfertilized eggs amongst vegetation. Relative fecundity can range 37490–163000 eggs/kg (Gurbuz et al. 2017), and increases with total length and age. This species can typically live for around 20 years in the wild (Vilizzi and Copp 2017).


Means of Introduction: There is some question as to when and where Common Carp was first introduced into the United States. DeKay (1842) reported that the species was first brought into the United States from France by Henry Robinson of Orange County, New York in 1831 and 1832. In a letter to DeKay, Robinson detailed that he kept the fish in ponds and for several years released one to two dozen carp during the spring in the Hudson River near his residence, thereby creating a commercial fishery for the species. S. F. Baird of the U.S. Fish Commission examined fish taken from the Hudson River, as well as area fish then being sold on the New York markets, and reported that they were goldfish or goldfish hybrids and not true Common Carp (Redding 1884; Cole 1905). Whitworth (1996) cited early literature indicating Common Carp had been introduced into Connecticut as early as the 1840s; however, the positive identity of the species is questioned. Smith (1896) reported that Common Carp first appeared in the United States in 1872 when J. A. Poppe of Sonoma, California, imported five specimens from Germany and propagated them in private ponds for commercial purposes, mainly distributing them to applicants as a food fish (Smith 1896; Lampman 1946). In 1877, the U.S. Fish Commission imported Common Carp from Germany and for the next two decades the agency began stocking and distributing the species as food fish throughout much of the United States and its territories (Smiley 1886; Smith 1896; Cole 1905). State fish commissions also were commonly involved in distributing the species (e.g., Johnson and Becker 1980). Records from the early 1880s indicate that Common Carp stocked in farm ponds frequently escaped into open waters as a result of dam breaks or flood events (Smiley 1886). By 1885, the U.S. Fish Commission was actively stocking lakes and rivers throughout the country. Often the fish were released from railroad tank cars at bridge crossing directly into streams (e.g., McDonald 1886). As a result of subsequent population growth and dispersal, Common Carp spread even further. More recently introductions of Common Carp have resulted because of the use of juvenile carp as bait fish (e.g., Swift et al. 1977). Various unusual genetic strains of Common Carp have been introduced into open waters of the United States. In addition to the normal scaled carp, the U.S. Fish Commission distributed both mirror carp and leather carp varieties in the late 1800s (Smiley 1886; Cole 1905). Colorful varieties of Common Carp (i.e., nishikigoi or koi) are kept as pets in garden ponds and some have been introduced to ponds and public water bodies (Balon 1995). However, only a small percentage of Common Carp records in U.S. open waters are based on koi. Another cultured variety occasionally found in open waters is the Israeli carp (Robison and Buchanan 1988). Their presence in South Florida is believed to be the result of released bait with this species as a contaminant.


Status: Recorded in all states except Alaska. In their summary table, Bailey and Smith (1981) indicated that Cyprinus carpio is widely distributed in the Great Lakes basin. Common Carp is only established in the Florida panhandle and does not appear to be established in South Florida. Populations are declining in the Upper Mississippi River system, potentially due to predator suppression, boom-bust population dynamics, outbreaks of cyprinid herpesviruses (CyHV-1 and CyHV-3), or resource exhaustion (Gibson-Reinemer et al. 2017).

Great Lakes:  Widespread, with populations reproducing and overwintering at self-sustaining levels in all five Great Lakes.


Great Lakes Impacts:
Summary of species impacts derived from literature review. Click on an icon to find out more...

EnvironmentalSocioeconomicBeneficial



Cyprinus carpio has a high environmental impact in the Great Lakes.

Realized:

The Common Carp is regarded as a pest fish in part because of its widespread abundance. Common Carp may destroy aquatic macrophytes directly by uprooting or consuming plants (Lee et al. 1980 et seq.), or indirectly by increasing turbidity, thereby reducing light for photosynthesis. This is accomplished by dislodging plants and rooting around in the substrate, thereby deteriorating habitat for species that require vegetation and clean water (Cole 1905; Cahoon 1953; Bellrichard 1996; Laird and Page 1996).

The role of Common Carp as an ecosystem engineer is well documented. For instance, following the installation of a carp barrier at Cootes Paradise Marsh (Lake Ontario), average turbidity was reduced by 40% in open water and 60% in vegetated areas, although further implications for plants and wildlife were difficult to assess due to variation in environmental conditions (Lougheed et al. 2004). Dentler (1993) found that Common Carp feeding behavior can destroy rooted aquatic plants which typically provide habitat for native fish species and food for waterfowl. One study analyzed the relationship between Common Carp biomass, vegetative cover, and waterfowl abundance over time in a shallow inland lake in Illinois (Bajer et al. 2009). The authors found that small densities of Common Carp (<30 kg/ha) did not have significant effects on vegetation or waterfowl, but a subsequent increase to over 250 kg/ha was strongly correlated with a decrease in vegetative cover from its original value of 94% to just 17% (Bajer et al. 2009). Furthermore, waterfowl activity dropped to ~10% of its original value. The authors suggested a threshold of 100 kg/ha past which Common Carp exert extensive ecological damage to shallow lakes (Bajer et al. 2009). In California, Common Carp have been implicated in a decline in water clarity in Clear Lake, Lake County, and in the gradual disappearance of native fishes (Moyle 1976).

A great deal of the Common Carp’s environmental impact is thought to come from indirect effects on habitat and the environment. For instance, in Mexico, populations of a native crayfish (Cambarellus montezumae) notably decreased with increasing carp density (Hinojosa-Garro and Zambrano 2004). However, further analysis indicated that Common Carp was not consuming the crayfish; rather, the destruction and depletion of crayfish habitat by Common Carp, particularly of algal species and macrophytes, were deemed to be the major mechanism of crayfish decline (Hinojosa-Garro and Zambrano 2004).

Miller and Crowl (2006) executed research in a eutrophic lake involving in situ observations of Common Carp impact through the use of cages and exclosures. They documented both direct and indirect effects of Common Carp on overall species composition, abundance, and plant species diversity. Common Carp also appeared to have indirect effects on macroinvertebrate community composition (Miller and Crowl 2006). A similar experiment set up enclosures within experimental ponds and noted that higher biomasses of Common Carp were positively related to phosphorus level, turbidity, and zooplankton biomass and negatively related to abundance of macroinvertebrates and macrophytes (Parkos et al. 2003). In comparison, channel catfish (Ictalurus punctatus), a native benthivore, affected phosphorus concentration and zooplankton communities, but had no significant effect on turbidity, macroinvertebrates, macrophytes, or suspended solids (Parkos et al. 2003).

In a biomanipulative experiment, Schrage and Downing (2004) removed >75% of the Common Carp population in Ventura Marsh, IA. In comparison to the adjacent reference site, they found that the removal of Common Carp had cascading effects, including an improvement in water quality related to decreased suspended solid and phytoplankton biomass. Within a few weeks, the authors noted an increase in Daphnia sp. and Ceriodaphnia sp. biomass as well as macrophyte diversity and density. The major limiting factor on maximum phytoplankton biomass appeared to switch from phosphorus abundance to zooplankton abundance, as suspended inorganic sediment settled to the bottom (Schrage and Downing 2004). Similarly, the eradication of Common Carp from three tributaries of the Bowman-Haley reservoir, North Dakota resulted in upwards of a 50-fold increase in chironomid densities (Bonneau and Scarnecchia 2015).

Common Carp has also been experimentally added to freshwater coastal wetland sites (Delta Marsh, Manitoba, Canada) at densities of 150, 300, 600, and 1200 kg•ha-1 (Badiou and Goldsborough 2010). The authors found that density of Common Carp was positively related to nutrient concentrations in the water column, suspended solids, and chlorophyll a concentrations. Furthermore, carp density was negatively related to dissolved oxygen concentrations, photic depth, and submersed macrophyte density (Badiou and Goldsborough 2010). These findings support the hypothesis that Common Carp may facilitate phytoplankton growth via increased nutrient loading in the water, effectively mimicking the effects of eutrophication (Badiou and Goldsborough 2015). Nevertheless, significant reduction in submersed macrophyte biomass was not observed, possibly because turbidity was relatively limited and the euphotic zone continued to span the entire water column at all carp densities (Badiou and Goldsborough 2010). Their results also suggested that suspension of solids increases as the colonized water body decreases in size, possibly due to a limited prey populations and increased forage activity by Common Carp. In this system, Common Carp populations were estimated to resuspend 37 to 361 kg of sediment per day, relative to pre-stocked conditions (Badiou and Goldsborough 2010). In Kohlman Lake, Minnesota areas with Common Carp increased sediment mixing depths 2.5 times greater (13.0 +/- 3.7 cm) than areas without Common Carp. The increase in mixing depth increased the amount of mobile phosphorus available for release by 55–92%, which could negatively affect the efficacy of nutrient management programs (Huser et al. 2016). Common Carp have been shown to alter bottom-up and top-down processes within freshwater ecosystems. Bottom-up processes are impacted by the alteration of nutrient flows and resuspension of turbidity and top-down are impacted by predation of zooplankton and benthic invertebrates. It can also decrease foraging efficiency of native species by impairing water quality (Weber and Brown 2009).

There is evidence that Common Carp prey on the eggs of other fish species (Moyle 1976; Taylor et al. 1984; Miller and Beckman 1996). For this reason, it may be responsible for the decline of the Razorback Sucker (Xyrauchen texanus) in the Colorado River basin (Taylor et al. 1984). In another case, Miller and Beckman (1996) documented White Sturgeon (Acipenser transmontanus) eggs in the stomachs of Common Carp in the Columbia River. In their review of the literature, Richardson et al. (1995) concluded that Common Carp has had notable adverse effects on biological systems, including the destruction of vegetated breeding habitats used by both fishes and birds. According to McCarraher and Gregory (1970), in 1894 it was documented that endemic Sacramento Perch (Archoplites interruptus) were becoming more scarce because Common Carp was destroying their spawning grounds.

Potential:

Laird and Page (1996) stated that Common Carp may compete with ecologically similar species such as carpsuckers and buffalos. Because this species has been present in many areas since initial surveys were completed, its impacts on many of the native fishes are difficult to determine.

Cyprinus carpio has hybridized with goldfish (Carassius auratus) and, in Europe, with the locally native crucian carp (Carassius carassius). However, Crucian x Common Carp hybrids were found in just 3 of 10 populations in which the two species geographically overlapped (Taylor and Mahon 1977; Hänfling et al. 2005).  Cyprinus carpio also hybridized with prussian carp (Carassius gibelio) (Balashov et al. 2017).

The destruction of macrophyte beds in two Spanish lakes by Common Carp negatively impacted the abundance of numerous waterfowl, including ducks, grebes, and flamingos. However, Common Carp served as a food source for gray herons (Ardea cinerea), bolstering their populations (Mediterreana Maceda-Veiga et al. 2017).

A variety of viruses currently infect US populations of Common Carp, including Koi herpesvirus (KHV; cyprinid herpesvirus-3; CyHV-3), carp oedema virus (CEV), Spring viremia of carp (SVC). Infections often lead to mass mortality events in Common Carp (Lovy et al. 2018; Thresher et al. 2018). Infection so far has been limited to Common Carp, but there is uncertainty if any of these viruses could infect new hosts. However,  McColl et al. (2017) found no sign of infection or toxic effects of KHV in a wide range of non-target species, including fish, crustacean, amphibian, reptile, and mammals.

Current research on the socio-economic impact of Cyprinus carpio in the Great Lakes is inadequate to support proper assessment.

Realized:

Once established in a waterbody, Common Carp is difficult and expensive to eliminate (e.g., Cahoon 1953). In a study of 129 lakes in Iowa, a negative relationship was discovered between Common Carp and sportfish abundance: (Bluegill (Lepomis macrochirus), Largemouth Bass (Micropterus salmoides), Black Crappie (Pomoxis nigromaculatus), and White Crappie (P. annularis)) (Jackson et al. 2010). This relationship could be due to the poor water quality (e.g., high nutrient levels and low water clarity), which was also associated with high Common Carp abundance; however, Common Carp’s role in the decline of the sportfish populations was not conclusively determined (Jackson et al. 2010).

Common Carp is fished commercially in the Great Lakes (Brown et al. 1999; Dann and Schroeder 2003). However, a recent study of contaminant levels in Lake St. Clair and the St. Clair River indicated that while most carp were below the general human consumption guidelines for mercury content, high PCB levels are of concern for both sensitive and general populations, especially in medium- to large-size fish (Gewurtz et al. 2010). Common carp have been found to contain high concentrations of PCB that are unsafe for consumption, especially within areas of concern in the Great Lakes (Bhavsar et al. 2018).

Anecdotally, Common Carp is widely considered to be a low value “trash” fish in the Great Lakes region. Coupled with real and perceived high contaminant burden, Common Carp is generally considered to be of low or even negative value to sport fishers. Peer-reviewed documentation of this aspect of the socio-economic impact was not able to be found.

Cyprinus carpio has a moderate beneficial effect in the Great Lakes.

Realized:

Common Carp has high lipid content and has been used to test contamination levels in the Great Lakes for comparison with human consumption guidelines (Gewurtz et al. 2010; Pérez-Fuentetaja et al. 2010).

Furthermore, Common Carp is fished commercially in the Great Lakes by both Canada and U.S. (Becker 1983; Brown et al. 1999; Dann and Schroeder 2003). It is also important as ornamental/aquarium fish, particularly if subspecies are considered (koi) (Rixon et al. 2005). It is a popular sport fish in parts of the U.S. According to Scott and Crossman (1973), the recreational pursuit of Common Carp was not considered common in Canadian waters historically, although it has been gaining popularity among anglers and in the tourism fisheries and fish markets in the Great Lakes region. Becker (1983) also described the growing presence of Common Carp in many branches of Wisconsin’s recreational and commercial fisheries.

Common Carp was shown to be an important seed dispersal vector for aquatic plants. However, it may also disperse nonindigenous plants (VonBank et al. 2018). Common Carp may serve as a food source for other organisms, as it was the primary fish consumed by North American river otters (Lontra canadensis) in central Illinois (Fretueg et al. 2015).

Potential:

Common Carp is commonly used in aquaculture in Mexico and Central America, South America, and Eurasia (FAO 2005). Global aquaculture production of Common Carp increased 10.4% per year between 1993 and 2002. At over 33 million tons in 2002, it made up nearly 14% of the global freshwater aquaculture production (FAO 2005). Also, fish oil harvested from Common Carp is a potential feedstock for biodiesel production (Fadhil et al. 2015).


Management: Regulations (pertaining to the Great Lakes region)

An individual must not use live Common Carp as bait in Indiana (312 IAC 9-6-8). This species is regulated in Minnesota and is legal to possess, sell, buy, and transport, but it may not be introduced into a free-living state, such as being released or planted in public waters (§ 84D.07). It is regulated in New York and cannot be knowingly introduced into a free-living state (6 NYCRR Part 575). In Pennsylvania it is unlawful for a person to use or possess this species as bait fish while fishing (58 Pa. Code § 63.44). It is a restricted species in Wisconsin, where there is a ban on the transport, transfer and introduction of this species, but possession is allowed (Chapter NR 40, Wis. Adm. Code). While not listed by name, in Ohio it is illegal for any person to possess, import or sell exotic species of fish (including Cyprinus carpio) or hybrids thereof for introduction or to release into any body of water that is connected to or otherwise drains into a flowing stream or other body of water that would allow egress of the fish into public waters, or waters of the state, without first having obtained permission (OAC Chapter 1501:31-19). In Canada, the use or possession of Cyprinus spp. as live bait is prohibited (SOR/93-55). It is illegal to bring any live fish into Ontario for use as bait (SOR/2007-237). In Quebec, both the use of this species as bait and the sale of dead fish of this species or its hybrids is prohibited (CQLR c C-61.1, r7 SOR/90-214).

Control

Biological

Northern Pike (Esox lucius) have been used as a biological tool to control Common Carp recruitment in the Sandhill lakes in Nebraska (Paukert et al. 2003). Bluegill (Lepomis macrochirus) are also an effective predator of juvenile carp and eggs (Bajer et al. 2012).

Inducible Fatality Genes (IFG) involve breeding carp with a fatal genetic weakness to a trigger substance, such as zinc. The fatal gene technology appears to be a potentially viable and long-term strategy for the environmentally benign control of carp (Koehn et al. 2000).

A variety of viruses currently infect US populations of Common Carp, often leading to mass mortality events (Lovy et al. 2018; Thresher et al. 2018). Both Koi herpesvirus (KHV; cyprinid herpesvirus-3; CyHV-3) and carp oedema virus (CEV) coinfect wild Common Carp populations in the midwest USA (Padhi et al. 2019). A 0.3–0.5% salt treatment may mitigate the negative effects of the disease (Stevens et al. 2018). Some of these viruses have been proposed as control mechanisms for Common Carp invasions. Spring viremia of carp (SVC) has been suggested as a control of Common Carp in Australia. However, releasing water-borne viral control agents would be controversial (Koehn et al. 2000). Koi herpesvirus is another potential biocontrol agent, but concerns exist that its purposeful release could result in re-colonization and recruitment boom of immune and virus resistant carp (Kopf et al. 2019; McColl and Sunarto 2020). Despite this, McColl et al. (2017) found no sign of infection or toxic effects of KHV in a wide range of non-target species, including fish, crustacean, amphibian, reptile, and mammals.

Physical

Barriers including electric, bubble curtain, and sonic have been used to exclude carp from industrial cooling intake structures (Koehn et al. 2000). Harvesting is only effective if carp are of importance by fisheries and anglers but success may be limited in large interconnected systems due to compensatory mortality and interbasin movement (Weber et al. 2016). Even if carp are beneficial for harvest, this method is one of the least effective methods available (Linfield 1980; Vacha 1998; Wedekind et al. 2001; in Arlinghaus & Mehner 2003; Koehn et al. 2000). Removal projects have included mechanical harvest by netting (Fritz 1987; Pinto et al. 2005), water level manipulation to disrupt spawning (Summerfelt 1999) and exclusion from spawning habitat with physical barriers (Lougheed and Chow-Fraser 2001). When possible, carp can be excluded from an area and then kept out through sorting of fish, which has been done since 1997 at the Cootes Paradise Marsh in Hamilton, Ontario (Lougheed et al. 2004).  Although labor intensive, this method is effective at keeping carp from returning to the marsh. Passive physical barriers are also a viable option to exclude Common Carp from specific habitats (such as spawning sites), and do not require manual sorting. These physical barriers can be based on differences in morphology across Common Carp (which tend to be wider) and native species, whereby mesh or vertical bars exclude Common Carp (Rahel and McLaughlin 2018). Other non-physical barriers based on sensory capabilities include carbon dioxide, which produces an avoidance behavior (Bzonek et al. 2021).

Common Carp display jumping behavior when trying to escape entrapment. The Williams cage exploits this behavior by selectively removing the jumping carp from other fish (Stuart et al. 2011). Tests of the Williams cage in Australia proved to be extremely successful. Over the three year testing, the Williams cage successfully separated 88% of adult Common Carp and allowed 99.9% native species to pass through. The Williams cage is useful in controlling dispersal and abundance of Common Carp. However, partial exclusion methods (including the Williams cage and other separation traps) can sometimes improve spawning conditions for resident Common Carp and increase their population growth rate and decrease generation time, offsetting the success of the exclusion mechanism. Thus, multiple control methods (e.g., harvest, wetland draining, etc.) need to be implemented concurrently to maximize success (Conallin et al. 2016; Casjenette et al. 2018). Further, the use of a population dynamic model for Common Carp in Malheur Lake, Oregon indicated that no single active control method (commercial harvest of adults, juvenile trapping, and embryo electroshocking) could sufficiently decrease Common Carp biomass, but the combination of two or three methods could due to cumulative mortality at multiple life stages (Pearson et al. 2019).

Chemical

Rotenone is a widely used non-selective chemical used to eliminate Common Carp from a water body (Sorensen & Stacey 2004). Application of different pheromones such as migratory, alarm, and sex may be useful in the integrated management of carp (Sorensen & Stacey 2004).

Antimycin-impregnated baits have been used to target Common Carp (Rach et al. 1994). The bait pellets consisted of fish meal, a binding agent, antimycin and water. Doses of 10 mg antimycin/g bait caused low (19%) to high (74%) mortalities in fish feeding voluntarily on 50 g of the toxic bait in each of three earthen ponds (Clearwater et al. 2008).

Note: Check state/provincial and local regulations for the most up-to-date information regarding permits for control methods. Follow all label instructions.


Remarks: Balon (1995) reviewed the origin and history of domestication of Common Carp in Europe and elsewhere. Several agents of the U.S. Fish Commission documented the early years of Common Carp propagation and stocking in the United States (e.g., Smiley 1886; Smith 1896; Cole 1905). Although this species was popular in the early 1870s as a food fish, Common Carp fell into wide disfavor soon after and is now considered a nuisance fish because of its abundance and detrimental effects on aquatic habitats. Trautman (1981) found Common Carp most abundant in streams enriched with sewage or with substantial runoff from agricultural land, but he reported it to be rare or absent in clear, cold waters, and streams of high gradient. Pflieger (1997) reported that the total weight and value of Common Carp taken by commercial fishermen in Missouri exceeded that of any other fish. Hartel et al. (1996) noted that more than 20,000 Common Carp were killed by a bacterial disease over a short period of time in the Merrimack River in the late 1970s. Because Common Carp have a higher salinity tolerance than most freshwater fishes, Swift et al. (1977) hypothesized that it may be spreading from one coastal stream to another through fresh or nearly fresh coastal waters in the Gulf area during periods of heavy rainfall and run-off, periods when salinities are greatly reduced. DeVaney et al. (2009) performed ecological niche modeling to examine the invasion potential for Common Carp and three other invasive cyprinids (Grass Carp Ctenopharyngodon idella, Black Carp Mylopharyngodon piceus, and Tench Tinca tinca). The majority of the areas where Common Carp have been collected, stocked, or have become established had a high predicted ecological suitability for this species.

Voucher specimens: Alabama (UMMZ 103508, 115003, TU 48856, 51966, 130781), Arizona (TU 74792, 78489, 79742), Arkansas (TU 2194, 2204, 44759), Colorado (TU 47337), Florida (TU 22858, 22879, 23654, 34833), Georgia (UGAMNH), Illinois (TU 9944, 125802, 125825), Indiana (TU 19372, 101143), Kansas (TU 42664, 42681), Kentucky (TU 66289), Louisiana (TU 6281, 9202, 15805, 16781), Michigan (TU 15007), Mississippi (TU 32974, 57121, 69483, 85130), Missouri (TU 53843, 54574, 74298), Nevada (TU 47257, 47266), New Jersey (TU 36738), New Mexico (TCWC 0059.01, TU 35686, 38871, 42637, 42656), New York (TCWC 0077.01, TU 36674), North Carolina (TU 29401), North Dakota (UMMZ 94756, 94757), Ohio (TU 3299), Oklahoma (TU 12021, 13790, 141667, 141686), Oregon (TU 121816), South Carolina (TU 145144), South Dakota (TU 58222), Tennessee (TU 33470), Texas (TCWC 1074.01, 07780.03, TU 15777, 21969, 21995, 35583, 35634), Utah (TU 43659, 99064, 99122, 99150), Wisconsin (TU 15748, 173824), many others.


References (click for full reference list)


Author: Nico, L., E. Maynard, P.J. Schofield, M. Cannister, J. Larson, A. Fusaro, M. Neilson, and A. Bartos


Contributing Agencies:
NOAA GLRI Logo


Revision Date: 4/5/2022


Peer Review Date: 4/5/2022


Citation for this information:
Nico, L., E. Maynard, P.J. Schofield, M. Cannister, J. Larson, A. Fusaro, M. Neilson, and A. Bartos, 2024, Cyprinus carpio Linnaeus, 1758: U.S. Geological Survey, Nonindigenous Aquatic Species Database, Gainesville, FL, and NOAA Great Lakes Aquatic Nonindigenous Species Information System, Ann Arbor, MI, https://nas.er.usgs.gov/queries/greatlakes/FactSheet.aspx?Species_ID=4&Potential=N&Type=0, Revision Date: 4/5/2022, Peer Review Date: 4/5/2022, Access Date: 4/23/2024

This information is preliminary or provisional and is subject to revision. It is being provided to meet the need for timely best science. The information has not received final approval by the U.S. Geological Survey (USGS) and is provided on the condition that neither the USGS nor the U.S. Government shall be held liable for any damages resulting from the authorized or unauthorized use of the information.